Lyme Borreliosis: Biology, Epidemiology and Control (CABI Publishing) [First ed.] 0851996329, 9780851996325, 9780851997551

Lyme borreliosis commonly known as lyme disease is now acknowledged as the most highly prevalent arthropod-borne human d

274 50 3MB

English Pages 480 [358] Year 2002

Report DMCA / Copyright

DOWNLOAD PDF FILE

Recommend Papers

Lyme Borreliosis: Biology, Epidemiology and Control (CABI Publishing) [First ed.]
 0851996329, 9780851996325, 9780851997551

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Lyme Borreliosis Biology, Epidemiology and Control

Lyme Borreliosis Biology, Epidemiology and Control Edited by

J. Gray Department of Environmental Resource Management University College Dublin Republic of Ireland

O. Kahl Institut für Angewandte Zoologie Freie Universität Berlin Germany

R.S. Lane Department of Environmental Science, Policy and Management University of California at Berkeley USA and

G. Stanek Institute of Hygiene University of Vienna Austria

CABI Publishing

CABI Publishing is a division of CAB International CABI Publishing CAB International Wallingford Oxon OX10 8DE UK Tel: +44 (0)1491 832111 Fax: +44 (0)1491 833508 E-mail: [email protected] Web site: www.cabi-publishing.org

CABI Publishing 10 E 40th Street Suite 3203 New York, NY 10016 USA Tel: +1 212 481 7018 Fax: +1 212 686 7993 E-mail: [email protected]

© CAB International 2002. All rights reserved. No part of this publication may be reproduced in any form or by any means, electronically, mechanically, by photocopying, recording or otherwise, without the prior permission of the copyright owners. A catalogue record for this book is available from the British Library, London, UK. A catalogue record for this book is available from the Library of Congress, Washington DC, USA. ISBN 0 85199 632 9

Typeset by Wyvern 21 Ltd, Bristol Printed and bound in the UK by Cromwell Press, Trowbridge

Contents

Contributors

vii

Preface

ix

1. History and Characteristics of Lyme Borreliosis G. Stanek, F. Strle, J. Gray and G.P. Wormser

1

2. Ecological Research on Borrelia burgdorferi sensu lato: Terminology and Some Methodological Pitfalls O. Kahl, L. Gern, L. Eisen and R.S. Lane

29

3. Molecular and Cellular Biology of Borrelia burgdorferi sensu lato S. Bergström, L. Noppa, Å. Gylfe and Y. Östberg

47

4. Vectors of Borrelia burgdorferi sensu lato L. Eisen and R.S. Lane

91

5. Borrelia burgdorferi sensu lato in the Vertebrate Host K. Kurtenbach, S.M. Schäfer, S. de Michelis, S. Etti and H.-S. Sewell

117

6. Ecology of Borrelia burgdorferi sensu lato in Europe L. Gern and P.-F. Humair

149

7. Ecology of Borrelia burgdorferi sensu lato in Russia E.I. Korenberg, N.B. Gorelova and Y.V. Kovalevskii

175

v

vi

Contents

8. Ecology of Borrelia burgdorferi sensu lato in Japan and East Asia K. Miyamoto and T. Masuzawa

201

9. Ecology of Borrelia burgdorferi sensu lato in North America J. Piesman

223

10. Epidemiology of Lyme Borreliosis D.T. Dennis and E.B. Hayes

251

11. Vaccination against Lyme Borreliosis E.B. Hayes and M.E. Schriefer

281

12. Environmental Management for Lyme Borreliosis Control K.C. Stafford and U. Kitron

301

Index

335

Contributors

S. Bergström, Department of Molecular Biology, Umeå University, SE-901 87 Umeå, Sweden. D.T. Dennis, Division of Vector-borne Infectious Diseases, National Center for Infectious Diseases, Centers for Disease Control and Prevention, Public Health Service, US Department of Health and Human Services, PO Box 2087, Fort Collins, CO 80522, USA. L. Eisen, Department of Environmental Science, Policy and Management, University of California, Berkeley, CA 94720, USA. S. Etti, Department of Infectious Disease Epidemiology, Imperial College of Science, Technology and Medicine, St Mary’s Campus, Norfolk Place, London W2 1PG, UK. L. Gern, Institut de Zoologie, University of Neuchâtel, Rue Emile-Argand 11, Case Postale 2, CH 2007 Neuchâtel 7, Switzerland. N.B. Gorelova, Gamaleya Research Institute for Epidemiology and Microbiology, Russian Academy of Medical Sciences, 18 Gamaleya Street, Moscow, 123098 Russia. J. Gray, Department of Environmental Resource Management, University College Dublin, Ireland. Å. Gylfe, Department of Molecular Biology, Umeå University, SE-901 87 Umeå, Sweden. E.B. Hayes, Division of Vector-borne Infectious Diseases, National Center for Infectious Diseases, Centers for Disease Control and Prevention, Public Health Service, US Department of Health and Human Services, PO Box 2087, Fort Collins, CO 80522, USA. P.-F. Humair, Institute of Zoology, University of Neuchâtel, Chantemerle 22, 2000 Neuchâtel, Switzerland. O. Kahl, Institute of Applied Zoology/Animal Ecology, Free University of Berlin, 12163 Berlin, Germany. Present address: Blackwell Verlag, Kurfürstendamm 57, 10707 Berlin, Germany. U. Kitron, Department of Veterinary Pathobiology, College of Veterinary Medicine, University of Illinois, 2001 S. Lincoln, Urbana, IL 61802, USA. vii

viii

Contributors

E.I. Korenberg, Gamaleya Research Institute for Epidemiology and Microbiology, Russian Academy of Medical Sciences, 18 Gamaleya Street, Moscow, 123098 Russia. Y.V. Kovalevskii, Gamaleya Research Institute for Epidemiology and Microbiology, Russian Academy of Medical Sciences, 18 Gamaleya Street, Moscow, 123098 Russia. K. Kurtenbach, Department of Infectious Disease Epidemiology, Imperial College of Science, Technology and Medicine, St Mary’s Campus, Norfolk Place, London W2 1PG, UK. R.S. Lane, Department of Environmental Science, Policy and Management, University of California, Berkeley, CA 94720, USA. T. Masuzawa, Department of Microbiology, School of Pharmaceutical Sciences, University of Shizuoka, 52-1 Yada, Shizuoka 422-8526, Japan. S. de Michelis, Department of Infectious Disease Epidemiology, Imperial College of Science, Technology and Medicine, St Mary’s Campus, Norfolk Place, London W2 1PG, UK. K. Miyamoto, Department of Parasitology, Asahikawa Medical College, Asahikawa, 0788510, Japan. L. Noppa, Department of Molecular Biology, Umeå University, SE-901 87 Umeå, Sweden. Y. Östberg, Department of Molecular Biology, Umeå University, SE-901 87 Umeå, Sweden. J. Piesman, Division of Vector-borne Infectious Diseases, National Center for Infectious Diseases, Centers for Disease Control and Prevention, Public Health Service, US Department of Health and Human Services, PO Box 2087, Fort Collins, CO 80522, USA. S.M. Schäfer, Department of Infectious Disease Epidemiology, Imperial College of Science, Technology and Medicine, St Mary’s Campus, Norfolk Place, London W2 1PG, UK. M.E. Schriefer, Division of Vector-borne Infectious Diseases, National Center for Infectious Diseases, Centers for Disease Control and Prevention, Public Health Service, US Department of Health and Human Services, PO Box 2087, Fort Collins, CO 80522, USA. H.-S. Sewell, Department of Infectious Disease Epidemiology, Imperial College of Science, Technology and Medicine, St Mary’s Campus, Norfolk Place, London W2 1PG, UK. K.C. Stafford, Connecticut Agricultural Experiment Station, PO Box 1106, New Haven, CT 06540, USA. G. Stanek, Clinical Institute of Hygiene and Medical Microbiology, University of Vienna, Austria. F. Strle, Department of Infectious Diseases, University Medical Centre, Ljubljana, Slovenia. G.P. Wormser, Division of Infectious Diseases, Department of Medicine, New York Medical College and Westchester Medical Center, Valhalla, New York, USA.

Preface

Lyme borreliosis, often referred to as Lyme disease, emerged during the mid1970s when an unusual form of juvenile arthritis was observed in the small town of Lyme, Connecticut, USA. Since then a wealth of literature on the subject has been published and eight international conferences as well as several smallerscale specialist or local meetings have taken place. Most publications, including several textbooks, have concentrated on clinical aspects, which reflects the importance of the disease in addition to the well-documented difficulties associated with diagnosis and treatment. The complex ecology of the causative agent, Borrelia burgdorferi, has also attracted much attention. Awareness of this complexity increased markedly when it became evident during the early 1990s that B. burgdorferi occurs as several distinct genospecies, referred to collectively as B. burgdorferi sensu lato, with apparently differing host preferences and pathological significance. In fact, most recent advances in our knowledge of Lyme borreliosis have concerned aspects of the basic biology and ecology of the causative spirochaetes and their tick vectors, leading to the development of vaccines and ecological preventive measures. However, the literature on these topics is widely scattered in research papers, reviews and meeting proceedings, and there is now a need for such information to be made available in book form. In this book, the first chapter outlines historical and clinical aspects including descriptions of the disease in humans and domestic animals, and of diagnosis and treatment. The second chapter deals with ecological methods and terminology and Chapters 3–5 describe the biology of the spirochaetes and their behaviour in vectors and vertebrates. The next four chapters (6–9) concern the ecology of B. burgdorferi s.l. in Europe, Russia, Southeast Asia and North America, and the last three chapters (10–12) deal with the application of the biological ix

x

Preface

and ecological attributes of the pathogens to disease epidemiology, vaccine development and the ecological management of Lyme borreliosis. The intended readership includes microbiologists, immunologists, zoologists, ecologists, epidemiologists, specialist clinicians, public health workers, specialist veterinarians, teachers, research students and all those interested in zoonotic or vector-borne infectious diseases. It is intended that the book will not only inform but will also stimulate much needed further research on the complex biology of B. burgdorferi s.l.

Acknowledgements We are especially grateful to Janet Robertson for her editorial assistance. Thanks are also due to the following authorities in the subject who gave generously of their time and expertise as reviewers: Lise Gern, Thomas Jaenson, Andreas Krause, Klaus Kurtenbach, Michael Levin, Jim Oliver, Eva Olsson, Joe Piesman, Danièle Postic, Sarah Randolph, Janet Robertson, Markus Simon, Susan Shaw, Dan Sonenshine, Reinhardt Straubinger, Franc Strle and Gary Wormser. We are also grateful to Bernard Kaye for creating some of the figures and for his assistance in the formatting of most of them.

History and Characteristics of Lyme Borreliosis

1

Gerold Stanek,1 Franc Strle,2 Jeremy Gray3 and Gary P. Wormser4 1

Clinical Institute of Hygiene and Medical Microbiology, University of Vienna, Austria; 2Department of Infectious Diseases, University Medical Centre, Ljubljana, Slovenia; 3Department of Environmental Resource Management, University College Dublin, Ireland; 4Division of Infectious Diseases, Department of Medicine, New York Medical College and Westchester Medical Center, Valhalla, New York, USA

History Before the first description of Lyme disease When the first cases were observed in Connecticut, USA, of a condition now known as Lyme disease (or Lyme borreliosis) (Steere et al., 1977a), it was 92 years after the first description in Europe of a chronic skin disorder (Buchwald, 1883) that was later named acrodermatitis chronica atrophicans (ACA) (Herxheimer and Hartmann, 1902). Thus, ixodid tick-borne borreliosis was present in Europe and emerged into clinical awareness at the end of the 19th century. In 1909, Afzelius presented a patient to the Swedish Dermatological Society in Stockholm with a slowly expanding skin lesion, which he called erythema migrans (EM) (Afzelius, 1910). A few years later the dermatologist Lipschütz from Vienna described a similar lesion in a 29-year-old woman, which expanded over 7 months from the back of the knee to the entire back and the neck (Lipschütz, 1913). Lipschütz was already aware of the transmission of some agent by the bite of a tick and he suggested that ‘. . . attention should be directed towards microscopic/bacteriologic investigations of the intestinal tract and of the salivary gland secretions of the tick’ (Lipschütz, 1923). However, no one followed his suggestions in Europe. The histological appearance of another skin manifestation of the disease, today called borrelial lymphocytoma, was first described in 1911 (Burckhardt, 1911). An extended description and definition of solitary lymphocytoma benigna cutis based on about 150 cases was given by Bäferstedt (1943). A tick bite was also the link to what is considered the first description of borrelial meningo-polyradiculoneuritis by the French neurologists Garin and © 2002 CAB International. Lyme Borreliosis: Biology, Epidemiology and Control (eds J. Gray, O. Kahl, R.S. Lane and G. Stanek)

1

2

G. Stanek et al.

Bujadoux (1922). The case report was an exciting story that elucidated the dramatic changes that may accompany borrelial infection of the central nervous system and/or the nerve roots. A relationship between tick bite, EM and disorders of the nervous system was first considered by Hellerström (1930). Bannwarth from Munich performed a careful analysis of several patients with chronic lymphocytic meningitis with the clinical syndrome of neuralgia and neuritis. He identified three groups, one with intense radicular pain, another with facial palsy and the third with chronic lymphocytic meningitis and cerebral symptoms such as severe headache and vomiting (Bannwarth, 1941). From today’s viewpoint it is evident that Bannwarth’s valuable work led to incorrect conclusions. He missed the relationship between tick bite and skin and nervous system manifestations and thought that the conditions were due to a rheumaticallergic cause. Despite this, Bannwarth’s name is now synonymous with meningopolyradiculoneuritis, the full picture of neuroborreliosis. Following the introduction of antibiotics after the Second World War, large studies were undertaken on the skin manifestations ACA and EM. A Swedish physician treated 57 patients with ACA with antibiotics and followed them for 2 years. He reported complete recovery and considerable improvement in about 70% (Thyresson, 1949). Successful antibiotic treatment stimulated the search for a causative agent of ACA and also renewed interest in the clinical course and pathogenicity of this chronic skin disease. More than 200 patients in Germany were evaluated in detail and associations between EM and lymphocytomas described by Hauser (1955). In a spectacular inoculation experiment, skin from an ACA lesion was successfully transplanted to the investigator and to three colleagues. An erythema-like rash developed on the skin of the volunteers around the borders of the inoculum; all skin changes disappeared after antibiotic therapy (Götz, 1954, 1955). Among the many studies performed at this time, one observation should be mentioned. Studies of ACA patients showed that peripheral neuropathy of the distal type was part of the clinical picture in a considerable proportion of patients and that the condition responded well to penicillin (Hopf, 1966). Electron microscope treatment studies were performed by Hollström (1958) in which it was reported that the EM disappeared within 2 weeks on average after the initiation of low-dose penicillin therapy. It was evident that an agent sensitive to antibiotics must play a role in the aetiology of these disorders. This was also suggested by observations of neurological disorders after EM and tick bite (Schaltenbrand, 1967). The most common neurological signs and symptoms after tick bites were painful radiculomeningitis, neuritis, myelitis and cranial nerve palsies that developed within a few weeks to 8 months (Sälde, 1946), and severe pain constituted the hallmark of the disease that lasted for 3 weeks on average before neurological deficit occurred (Hörstrup and Ackermann, 1973). In 1974, Weber reported on a patient who developed meningitis after EM, who responded very well to intravenous penicillin (Weber, 1974). Weber got very close to the ‘real’ aetiologic agent as he systematically and logically ruled out viruses and many tick-transmitted bacteria known at that time, and mentioned borreliae as possible causative

History and Characteristics of Lyme Borreliosis

3

agents. Unfortunately, he could not follow this track, and the enigmatic aetiology of these tick-borne clinical conditions persisted until the early 1980s.

A new clinical entity In October 1975, two mothers from Lyme, Connecticut, reported to the Connecticut State Health Department that a number of children who lived close together were thought to have juvenile rheumatoid arthritis. A surveillance system was organized in order to identify all children with inflammatory joint disease in the communities of Old Lyme, Lyme and East Haddam, Connecticut. The arthritis cases found were of a similar type, characterized by brief but recurrent attacks of swelling and pain in a few large joints, especially the knee, over a period of years. Parents and neighbours of the affected children were found to have a similar type of arthritis. Moreover, the frequency of this type of arthritis was 100 times greater than juvenile rheumatoid arthritis. The majority of patients noted the onset of arthritis in summer or early autumn. Furthermore, a certain proportion of patients noted a skin rash approximately 4 weeks before the onset of arthritis. This skin lesion was found to be compatible with EM, described in Europe many decades before (Afzelius, 1910; Lipschütz, 1913). However, the disease in Lyme residents was thought to be a previously unrecognized clinical entity and was called Lyme arthritis (Steere et al., 1977a,b). In the following years, the Yale investigators identified patients with the skin lesion EM and followed the patients prospectively. Many of them subsequently developed arthritis, and some also acquired neurological and cardiac abnormalities. Thus, a complex, multi-system disorder was recognized and the name was changed from Lyme arthritis to Lyme disease. Furthermore, the vector of the disease was identified as the hard tick Ixodes scapularis (then called Ixodes dammini). In the following years, Lyme disease was found to be present along the northeastern coast of the USA and also in Wisconsin, California and Oregon; there was strong evidence that ixodid ticks were the vectors (Steere and Malawista, 1979). The agent was apparently a non-pyrogenic, penicillin-sensitive bacterium since penicillin treatment shortened the duration of EM and appeared to prevent subsequent arthritis (Steere et al., 1980). But the agent was still unknown.

Discovery of the Lyme disease agent The discovery of the Lyme disease agent came unexpectedly in conjunction with a tick and rickettsial survey of Long Island, New York (Burgdorfer et al., 1982). Unable to find the Rocky Mountain spotted fever agent in Dermacentor variabilis ticks, the researchers speculated that possibly other ticks such as I. scapularis (dammini), which were there in abundance, might carry the agent. No rickettsiae were found in the midguts of many dissected ticks but poorly stained and

4

G. Stanek et al.

rather long, irregularly coiled spirochaetes were observed. The bacteria were isolated and cultivated in a modified Kelly’s medium, which is now known as BSKII medium and the organisms were linked with Lyme disease after they were observed to react with sera of Lyme disease patients in immunofluorescence assays (Burgdorfer et al., 1982; Barbour, 1984). Shortly thereafter the spirochaete was isolated from blood, skin and cerebrospinal fluid of a few Lyme disease patients (Benach et al., 1983; Steere et al., 1983).

A worldwide tick-borne disease Very soon after the successful isolation of a new borrelia from hard ticks and patients in the USA, researchers in Europe isolated spirochaetes from ticks, and from skin and cerebrospinal fluid of patients in Sweden, Germany, Switzerland and Austria (Preac-Mursic et al., 1984; Asbrink and Hovmark, 1985; Stanek et al., 1985). With the methods available initially, the spirochaetes appeared to be the same as the USA isolates. The bacterium was identified as a new species among borreliae and named Borrelia burgdorferi (Johnson et al., 1984). Barbour and Schrumpf (1986) raised monoclonal antibodies to the outer surface proteins (Osp) A and B, and to flagellin proteins of the spirochaetes, which resulted in the detection of substantial differences in reactivity. An OspA typing system was developed in Munich (Wilske et al., 1986) and the resulting findings in conjunction with molecular biological studies (Baranton et al., 1998) provided evidence that two new Borrelia genospecies, Borrelia afzelii and Borrelia garinii, were also human pathogens. These genospecies occur in Europe and Asia but not in North America. The USA type was called B. burgdorferi sensu stricto (s.s.), and the overall term for the genospecies complex is B. burgdorferi sensu lato (s.l.). The discovery of a new tick-transmitted agent sparked off a spate of research in various branches of science, including human and veterinary medicine, molecular biology, and tick and reservoir host biology and ecology. Lyme disease, or Lyme borreliosis (LB) as it is more widely known, is now recognized as a tickborne spirochaetal infection with a wide geographical distribution in the northern hemisphere. However, the story is still incomplete and new findings continue to emerge, including the discovery of additional genospecies.

Scientific publications and international conferences on tick-borne diseases Up to mid-July 2001, the PubMed service of the National Library of Medicine listed 5592 scientific publications under the search words ‘Lyme disease’, and research is still proceeding vigorously. Besides these publications there are numerous others available in English and in other languages, and many theses and several books on LB have been written and published since the early 1980s. An international conference on LB has been held every 2 or 3 years since 1983,

History and Characteristics of Lyme Borreliosis

5

and the number of smaller international symposia occurring in different places in the world is increasing (Table 1.1). Topics have included clinical manifestations, treatment regimens, concepts on prophylaxis, vectors, the still poorly understood changes of the agent during its infectious cycle, and studies on pathogenicity of the disease. The European Union Concerted Action on Lyme Borreliosis (EUCALB) (Gray and Stanek, 1998) was the first continent-wide approach to attempt risk assessment in relation to exposure and to improve diagnostic criteria for both case management and the quality of epidemiological data.

Lyme borreliosis as a social phenomenon Contrary to the success in identifying the aetiology of multi-system LB and in introducing proper treatment, people in endemic areas of the USA and subsequently in Europe have exhibited a high degree of anxiety concerning this disease (Barbour and Fish, 1993). Lyme disease/borreliosis self-help groups have been formed with ‘tick-victims’ as their members. These activities appear to be acceptable or even truly helpful as long as they form an information desk that guides persons to trained physicians who follow the therapeutic concepts of evidence-based medicine. Problems arise when such groups develop their own pictures of the disease. Moreover, certain physicians have increased the fear of patients with their private opinion that long-term treatment (over many months) is obligatory in order to get rid of the pathogens hidden in the body. Other physicians have developed obscure diagnostic procedures such as identification of borreliae in the peripheral blood by dark-field microscopy and adhere to this approach although it has been discredited by other methods including nucleic acid amplification techniques. Some consequences of such activities were described in an article published in the New York Times of June 2001 (Grann, 2001). What follows is a brief review of the salient features of the disease, its diagnosis and treatment.

Clinical Features Three stages of LB have been described, but the disease itself does not necessarily develop in stages. Neuroborreliosis, for example, is not infrequently preceded by EM but neurological disorders may also develop without it. Chronic manifestations such as ACA usually develop unrecognized initially and are not preceded by EM. The concept of localized versus disseminated infection is open to question because up to 40% of US patients with a single EM have microbiological evidence of haematogenous dissemination. The main clinical manifestations of LB are listed in Table 1.2 and Table 1.3 provides a simplified outline of the differences between the clinical features in Europe/Eurasia and those in North America. In the following paragraphs the various clinical manifestations will be introduced briefly.

6

G. Stanek et al.

Table 1.1. International conferences on Lyme disease/Lyme borreliosis. Year

Location

1983

New Haven, Connecticut, USA Vienna, Austria Baden, Austria

1985 1987 1987 1990 1992 1993 1994 1994 1995

1995 1996 1997 1997 1998

1999 1999

2001 2001 2002

Conference

I International Conference on Lyme Disease II International Conference on Lyme Disease and Related Disorders 1st European Symposium on Lyme Borreliosis New York, USA III International Conference on Lyme Disease and Related Disorders Stockholm, IV International Conference on Sweden Lyme Borreliosis Arlington, USA V International Congress on Lyme Borreliosis London, UK 2nd European Symposium on Lyme Borreliosis Bologna, Italy VI International Conference on Lyme Borreliosis Shizuoka, 1st International Symposium on Japan Lyme Disease in Japan Portoroz, International Symposium on the Slovenia Therapy and Prophylaxis for Lyme Borreliosis Warsaw, WHO Workshop on Lyme Borreliosis Poland Diagnosis and Surveillance San Francisco, VII International Congress on USA Lyme Borreliosis Berlin, 4th International Potsdam Germany Symposium on Tick-borne Diseases Shizuoka, Japan 2nd International Symposium on Lyme Disease in Japan Vienna, Austria International Symposium on the Pathogenesis and Management of Tick-borne Diseases Berlin, Germany 5th International Potsdam Symposium on Tick-borne Diseases Munich, Germany VIII International Conference on Lyme Borreliosis and Other Emerging Tick-borne Diseases Berlin, Germany 6th International Potsdam Symposium on Tick-borne Diseases Ljubljana, International Symposium on Slovenia Tick-transmitted Diseases New York, USA IX International Conference on Lyme Borreliosis and Other Tick-borne Diseases

Publication Yale J. Biol. Med. 57 1984 Stanek et al., 1986 Stanek et al., 1989 Benach and Bosler, 1988 Sköldenberg and Stiernsted, 1991 Abstracts only Axford and Rees, 1994 Cevenini et al., 1994 Yanagihara and Masuzawa, 1994 Infection 24 (1996) 60–102 24 (1996) 170–212 Anon., 1995 Abstracts only Süss and Kahl, 1997 Yanagihara and Masuzawa, 1997 Wien klin. Wochenschr. 110 (1998) 847–910 111 (1999) 909–1004 Süss and Kahl, 1999 Abstracts only

Süss et al. (2002) Wien klin. Wochenschr., in press

History and Characteristics of Lyme Borreliosis

7

Table 1.2. Main clinical manifestations of Lyme borreliosis. Organ system

Clinical feature

Skin

Erythema (chronicum) migrans Erythema migrans multiple lesions Borrelial lymphocytoma Acrodermatitis chronica atrophicans

Nervous system

Meningo-radiculoneuritis (Bannwarth syndrome) Cranial nerve palsy Meningitis Meningoencephalitis Encephalomyelitis Cerebral vasculitis Peripheral neuropathy

Musculoskeletal

Arthritis Myositis

Heart

Carditis

Eye

Conjunctivitis, keratitis, endophthalmitis, panophthalmitis

Skin manifestations Erythema (chronicum) migrans EM is the most frequent clinical manifestation of LB. It occurs at any age and in both sexes. EM is most frequently diagnosed in late spring and early summer but can be seen in individual cases throughout the year. It develops at the site of a tick bite, though the bites are often not detected by patients. The incubation period ranges from a few days to more than 1 month, but is usually about 10 days. The initial EM is commonly a homogeneous lesion, which may either be bright red or turn bluish-red, and the homogeneous nature may not change in some patients until treatment or spontaneous healing. In cases of long duration, the lesion will fade or clear in the central part and a peripheral bright red erythematous ring will further expand, again until treatment or spontaneous healing. How the lesion expands depends partly on the site of the initial infection. The border of EM is usually demarcated. In adult European patients, constitutional symptoms accompanying EM (most often fatigue, malaise, arthralgia, myalgia and headache) are present in fewer than 50% and are regularly mild, while local symptoms such as itching, burning and/or mild local pain are reported by about 50% of patients (Strle et al., 1993; Logar et al., 1999). Constitutional symptoms are more frequent in the USA (up to 15% of cases). The different shapes that can be observed with EM make a definition of the skin lesion a challenging task. Generally, EM is an enlarging erythematous to

8

G. Stanek et al.

Table 1.3. Lyme borreliosis in Eurasia and North America. (Modified after Steere, 2001.) Organ system Skin Acute phase

Chronic phase

Nervous system Acute phase

Chronic phase

Heart Acute phase Chronic phase

Musculoskeletal Acute phase

Chronic phase

Europe/Eurasia (Borrelia afzelii, B. garinii, B. burgdorferi s.s.)

North America (B. burgdorferi s.s.)

Erythema (chronicum) migrans Erythema migrans faster spreadslower spreading, less intensely ing, more intensely inflamed, of inflamed, of longer duration; less briefer duration; frequent, possibly frequent haematogenous diswidespread haematogenous semination, possible regional or dissemination. contiguous spread to other sites. Borrelial lymphocytoma rarely and predominantly seen in children and scarcely in adults. Acrodermatitis chronica atrophicans ACA rarely reported. (ACA), caused primarily by B. afzelii. Circumscribed scleroderma and None reported. Lichen sclerosus et atrophicus; reports on the isolation of B. afzelii from lesional skin, improvement after antibiotic therapy in single cases. Meningo-polyradiculoneuritis with severe radicular pain, pleocytosis and intrathecal antibody production, caused primarily by B. garinii. Subtle sensory neuropathy often within areas affected by acrodermatitis. Severe encephalomyelitis, spasticity, cognitive abnormalities, marked intrathecal antibody production.

Meningitis, less prominent radiculoneuritis.

Subtle sensory polyneuropathy without acrodermatitis. Subtle encephalopathy, cognitive disturbance, slight intrathecal antibody production.

Atrioventricular block and subtle myocarditis. Dilated cardiomyopathy; isolation of B. burgdorferi s.l. from endomyocardial biopsies in only a single case.

Atrioventricular block and subtle myocarditis. None reported.

Less frequent oligoarticular arthritis, less intense joint inflammation.

More frequent oligoarticular arthritis, more intense joint inflammation. Treatment-resistant arthritis in about 10% of patients, probably due to autoimmune mechanism.

Persistent arthritis rare.

History and Characteristics of Lyme Borreliosis

9

bluish-red patch with an advancing border, with or without central clearing. However, there are conditions with which these lesions may be confused and which should be considered. Differential diagnosis comprises non-specific insectbite reactions, erythema annulare centrifugum, urticaria, contact dermatitis, folliculitis, vasculitis, granuloma annulare, tinea corporis and fixed drug eruption. The histology of EM-lesional skin is usually unimpressive. Histological findings of skin samples from the central and peripheral parts of EM show a superficial perivascular infiltrate with lymphocytes, histiocytes, a few plasma cells, eosinophils and neutrophils. Blood vessels may show endothelial swelling, and extravasation of erythrocytes is frequently found. Borrelial lymphocytoma Lymphocytoma is a term that denotes cutaneous proliferation of lymphoid cells without detectable extracutaneous involvement. Because of its benign course, the lesion has also been called lymphadenosis benigna cutis. Borrelial lymphocytoma was the term introduced by Weber et al. (1985) in order to specify the aetiology of the condition (Hovmark et al., 1986a). Borrelial lymphocytoma presents as a bluish-red tumour-like skin infiltrate, up to a few centimetres in diameter, with lymphoreticular proliferation in the dermis and/or subcutis. The lesion may develop several weeks to months after a tick bite. Predilection sites are the ear lobe, ear helix, nipple or scrotum in children, and nipple in adults. The lesion may be preceded by EM or occur simultaneously with it. Borrelial lymphocytoma heals spontaneously but can persist and even grow for several months (Strle et al., 1992). This skin manifestation is not seen in the USA. The histological picture of borrelial lymphocytoma is impressive. It presents either as a diffuse, dense dermal polyclonal lymphocytic infiltrate or a follicular lesion with germinal centres. Acrodermatitis chronica atrophicans ACA usually begins on the extensor sites of the extremities, most commonly on the lower leg with initial involvement of one foot. Besides this, the soles and palms, toes, fingers and knees can be involved, and, in rare cases, ACA may also develop on the trunk. The lesion may persist over a long period of time and expand on one extremity. With time, however, the lesions will usually become symmetric but in some cases may be asymmetric, restricted to one extremity. The first signs are localized, livid-red colour changes and a doughy swelling of the involved skin. The swelling may be pronounced, which might cause the patient to seek medical attention. ACA does not heal spontaneously. Within years of onset, fibrous thickening may develop over, or close to, joints and after years of progression the lesional skin will gradually become atrophic, hyperpigmentation and sometimes depigmentation will occur and the veins may appear prominent because of the thin and wrinkled, ‘cigarette paper-like’ skin. Chronic ulceration may develop and form the basis for squamous or basal cell carcinomas. Sclerotic changes of the skin may develop, resembling morphoea.

10

G. Stanek et al.

The borrelial aetiology of ACA was confirmed by culture and serology (Asbrink and Hovmark, 1985). B. afzelii is predominantly isolated from samples of lesional skin, but the other pathogenic genospecies have been isolated from cases of ACA (Picken et al., 1998). ACA rarely, if ever, occurs in the USA. Associated with long-standing ACA is peripheral neuropathy (Kristoferitsch et al., 1988) and joint manifestations such as partial or complete dislocations of the small joints of hands and feet, clawed toes, periosteal thickening, joint effusion and hyperpathia, and periarticular manifestations (Asbrink et al., 1986). Chronic sclerodermic skin manifestations B. afzelii strains have been isolated from the lesional skin of some cases of morphoea (circumscribed scleroderma) (Breier et al., 1999, 2001).

Neuroborreliosis Early neurological manifestations The full picture of neuroborreliosis is defined as meningo-polyradiculoneuritis, also named Bannwarth syndrome. All age groups may be affected, although painful radiculoneuritis has only very rarely been observed in children. Several studies have reported that the median age for Bannwarth syndrome is the fifth decade, ranging from 6 to 80 years. The incubation period for neuroborreliosis is on average 7 weeks in adults and 4 weeks in children. In adults, this disorder often presents as a triad, which includes radicular pain, peripheral pareses, most frequently facial palsy, and lymphoplasmocytic pleocytosis in the cerebrospinal fluid with a cell count of up to 1000 × 106 l–1. The majority of cases of early neuroborreliosis occur between July and November, with most cases observed in August (Kristoferitsch, 1989). Nevertheless, the disorder is diagnosed throughout the year. The clinical picture of Bannwarth syndrome begins with pain because of radiculoneuritis. The character of the pain varies, is usually severe, and patients describe it as burning, tearing and migrating. A characteristic of neuroborreliosis pain is its aggravation at night. Also remarkable is that this pain is refractory to normal dosages of analgesics. Furthermore, the skin at the site of the pain may be dysaesthetic and hyperpathic. The slightest pressure by contact of cloth, for example, may cause severe burning pain. Pain constituted the only symptom in up to 20% of patients. Because of its severity, about one-third of patients become depressed, agitated, restless, suffer from sleeplessness and are full of anxiety. Even autoaggressive behaviour has been observed. Patients with pain syndrome alone have been certified insane as a result of personality changes and misdiagnosis. In most cases, however, the onset of pareses, especially cranial nerve involvement 1 or 2 weeks after onset of pain, leads to the correct diagnosis. Before the discovery of the agent of LB and the introduction of specific antibiotic treatment, the syndrome was observed in its spontaneous course, which lasted for up

History and Characteristics of Lyme Borreliosis

11

to 25 weeks until complete recovery or recovery with incomplete resolution of nerve function occurred. Neuroborreliosis in children presents predominantly as acute facial palsy and/or acute aseptic meningitis. Bannwarth syndrome was observed in fewer than 4% of all cases studied, and mostly in older children (Christen et al., 1993). Chronic neuroborreliosis The major manifestations of chronic neuroborreliosis are encephalitis, radiculomyelitis, transverse myelitis, stroke-like disorders and cranial nerve deficits. The condition is very rare, and diagnosis has to fulfil strict diagnostic criteria with other diagnostic considerations including neurosyphilis, fungal meningoencephalitis and brain tumour (Wokke et al., 1987; Ackermann et al., 1988).

Joint manifestations In the USA, arthritis was the predominant manifestation that led to the discovery of this complex borrelial infection. Episodes of frank arthritis occurred in about 50% of patients who did not receive antibiotic treatment for EM (Steere et al., 1987). In Europe, the situation is different and arthritis is not one of the major complications. Lyme arthritis may occur within several months after an unrecognized primary infection and the resolution of an untreated EM. It occurs predominantly in the fourth decade and, if in childhood, older children are more often affected. The joints most frequently involved are the knee in about half of all cases, and the ankle, wrist, elbow and rarely smaller joints (Herzer, 1991). Lyme arthritis is caused by an intra-articular infection (Priem et al., 1998). A small number of bacteria are liable to provoke severe arthritis by inducing mechanisms (including the induction of cytokines and chemokines) that amplify the inflammatory response (Franz et al., 1999).

Cardiac manifestations Acute cardiac involvement is a rarely observed manifestation that may present as conduction or rhythm disturbance, and occasionally as acute myopericarditis or mild left ventricular dysfunction (Steere, 2001). The condition occurs mostly in the course of an EM or within a few weeks after onset of infection, and appears to be transient and self-limiting. In Europe, however, B. burgdorferi s.s. was isolated from the endo-myocardium of a patient suffering from longstanding dilated cardiomyopathy (Stanek et al., 1990).

12

G. Stanek et al.

Lyme Borreliosis in Domestic Animals So far this review has considered LB only in humans. LB is categorized as a zoonotic disease in humans because the infection is maintained in nature by animals; humans are infected accidentally and are dead-end hosts for B. burgdorferi s.l. The animal reservoirs of B. burgdorferi s.l. are many and varied (see Kurtenbach et al., Chapter 5) and the majority are wild small mammals and birds. In common with many animal reservoirs of zoonotic infections, these animals do not seem to show signs of disease as a result of infection with B. burgdorferi s.l., but other animals, particularly domestic animals, can become victims of this abnormal host/pathogen relationship and may show similar manifestations to those that occur in humans. Most data on LB in domestic animals concern dogs and horses, with isolated reports of infection in cattle, sheep and cats. Anti-borrelial antibodies have been detected in a wide range of domestic animal species, but it is probable that most infections are subclinical and self-limiting. It is also possible that infections are missed or misdiagnosed because of the non-specificity of clinical symptoms. Domestic animals do not show any pathognomic symptoms such as EM and there are no case definitions for LB in animals. There are also considerable difficulties involved in laboratory diagnosis. In contrast with diagnostic methods for human LB, little quality assurance has been applied to the laboratory diagnosis of animal LB. Furthermore, it should be borne in mind that the tick vectors of LB transmit other pathogens, which may confound both clinical and laboratory diagnosis of LB (Gray, 1999).

Lyme borreliosis in dogs Canine LB was first reported in the original LB focus and neighbouring regions in the USA (Lissman et al., 1984; Kornblatt et al., 1985). Typical signs are lameness combined with malaise. Fever may accompany the malaise, which manifests as listlessness and inappetance, and the lameness may be extreme claudication with localized painful joints or a general stiff gait without clear pain localization. Kidney involvement (Dambach et al., 1997) and heart block (Levy and Duray, 1988) have also been reported. Diagnosis of canine LB requires evidence of exposure to ticks and the presence of typical clinical signs together with antibodies to B. burgdorferi s.l. As with all LB serology, problems arise with specificity and cross-reactions with other spirochaetal organisms such as Treponema spp. Positive serology is not diagnostic on its own since most infections are evidently subclinical, though serosurveys have shown that prevalence of B. burgdorferi s.l. antibodies is usually higher in symptomatic dogs compared with healthy ones (Hovius, 1999). A positive response to antibiotics (e.g. tetracyclines) is also regarded as supportive of the diagnosis, but there is evidence of persistence of clinical signs despite treatment (Straubinger et al., 1998). Increasing use is being made of PCR to support

History and Characteristics of Lyme Borreliosis

13

diagnosis (Appel et al., 1993; Hovius et al., 1999) but definitive demonstration of the organism by culture is rarely successful. In Europe serosurveys suggest widespread infection (Hovius, 1999) and there are several reports where typical symptoms have been accompanied by elevated antibody titres (Font et al., 1992; McKenna et al., 1995; Overduin and van den Bogaard, 1997; Hovius et al., 2000). B. burgdorferi s.l. can be detected in symptomatic dogs by PCR (Hovius et al., 1999), but so far only one study has reported diagnosis supported by isolation and identification of B. burgdorferi s.l. (Liebisch and Liebisch, 1999). More recently B. afzelii was isolated from the blood of a dog with neurological abnormalities, but the dog was euthanased before any definite association between the infection and the symptoms could be established (Speck et al., 2001). Despite the difficulties associated with unequivocal diagnosis, clinical canine LB is considered sufficiently prevalent in the USA and Europe for the marketing of both inactivated whole-cell and recombinant vaccines (Wormser, 1999).

Lyme borreliosis in cats LB has not been reported in cats exposed to natural infection though anti-B. burgdorferi antibodies have been detected (Magnarelli et al., 1990; Oliver et al., 1999). In one study, experimental infections resulted in seroconversion but not clinical manifestations (Burgess, 1992), whereas, in another, clinical manifestations were reported that apparently resembled those in humans sufficiently closely for this feline model to be suggested for vaccine testing (Gibson et al., 1995). There are insufficient data for evaluation of the significance of LB in cats but the latter observation suggests that these animals may be affected where they are exposed to a heavy tick challenge.

Lyme borreliosis in horses Suspected equine LB was reported in endemic areas in the USA a few years after identification of B. burgdorferi as the causal agent of the human disease (Burgess et al., 1986; Burgess and Mattison, 1987; Magnarelli et al., 1988). Serosurveys have shown that the infection is widespread, and that seroprevalence may be as high as 45% in some areas (Magnarelli et al., 2000). Serosurveys conducted in Europe have also shown high seroprevalence in some countries (Käsbohrer and Schönberg, 1990; Gerhards and Wollanke, 1996; Egenvall et al., 2001; Müller et al., 2001). Clinical signs of LB in horses consist mainly of malaise, low grade fever, stiffness, laminitis and swollen joints, and there are also several reports of eye involvement such as uveitis (Burgess et al., 1986; Parker and White, 1992; Hahn et al., 1996), two of cardiac problems (Browning et al., 1993; Liebisch et al., 1999) and at least one of encephalitis (Burgess and Mattison, 1987). However, clinical cases

14

G. Stanek et al.

appear to be uncommon and, although some may be missed because of the nonspecific nature of symptoms, it is apparent that in the vast majority of cases infection is asymptomatic. No clinical signs arose even when infection was induced experimentally with infected nymphal ticks in immunosuppressed animals (Chang et al., 2000). There is no evidence that horses can be a source of infection for ticks. Despite increased awareness, improvement of serological diagnostics and the advent of molecular methods (Manion et al., 1998), unequivocal diagnosis of equine LB is still very difficult. It is only recently that modified Koch’s postulates have been satisfied in the diagnosis of equine LB, in this case involving the presence of appropriate clinical signs in four animals exposed to ticks in Germany, successful isolation of spirochaetes and their identification by PCR as B. burgdorferi s.l. (Liebisch et al., 1999). However, the PCR product was apparently not sequenced and the evident rapidity of the growth of the isolated spirochaete culture has led to some doubts about its identity. There is some interest in the development of vaccines against equine LB, particularly in the USA (Chang et al., 1999), but, in view of the diagnostic difficulties and uncertain risk of disease in horses, justification for this preventive measure may require further attention.

Lyme borreliosis in cattle Several cases of clinical LB in cattle reported in the USA in the late 1980s and early 1990s were diagnosed by association of symptoms such as lameness, weight loss and abortion, with positive serology, isolation of B. burgdorferi from colostrum, serum, urine, milk or synovial fluids, apparent response to tetracyclines and seroconversion in mice and a cat after exposure to bovine urine and milk, respectively (Burgess et al., 1987; Burgess, 1988; Post et al., 1988; Wells et al., 1993). The apparent infectivity of milk could have profound implications for human health, but, despite continued interest in this topic, no more evidence has been forthcoming and it must be concluded that milk from infected animals (or humans) is not infective. In fact, very few cases of LB in cattle have been diagnosed subsequently, though isolated reports appeared from Australia (Rothwell et al., 1989), Japan (Isogai et al., 1992) and the UK (Blowey et al., 1994). Doubts concerning the specificity of the serology have been reinforced by studies implicating Treponema spp. as the causal agent of B. burgdorferi seropositivity of cattle with digital dermatitis (Choi et al., 1997; Demirkan et al., 1999) and crossreactions with other spirochaetes have also been recorded (Rogers et al., 1999). When cattle were experimentally infected with three pathogenic genospecies of B. burgdorferi s.l. (B. afzelii, B. burgdorferi s.s. and B. garinii) no clinical symptoms resulted (Tuomi et al., 1998), though infection by tick bite might have produced a different result. It appears likely that B. burgdorferi s.l. is usually non-pathogenic in cattle. However, a recent report strongly suggests that at least two genospecies may be pathogenic under certain circumstances (Lischer et al., 2000). In this

History and Characteristics of Lyme Borreliosis

15

Swiss study of two dairy cows, the clinical signs included erythema on the udder, loss of weight, swollen legs, swollen joints, stiff gait, acute laminitis and oligoarthritis. In order to decrease the possibility of erroneous results arising through contamination, real-time PCR was used as a diagnostic tool and B. burgdorferi s.s. DNA was detected in synovial fluid and milk of one cow and B. afzelii DNA in synovial fluid of the other. LB in cattle is probably an infrequent disease that is difficult to diagnose due to the non-specific nature of symptoms and doubt surrounding the specificity of serology. The application of PCR technology as a diagnostic aid may result in more certain identification of cases in the future. Cattle are not regarded as reservoir hosts of B. burgdorferi s.l. (Gern et al., 1998) and the association of infected ticks with cattle pastures (Borko and BoleHribovsek, 1998) is probably due to the presence of other hosts in the same habitat (Gray et al., 1995).

Lyme borreliosis in sheep In the only report of possible LB in sheep, the suspected disease manifested in seropositive Norwegian lambs as pronounced lameness in one leg (Fridriksdottir et al., 1992). In the absence of other reports it must be concluded that B. burgdorferi s.l. infection of sheep rarely results in disease. There are several reports of the detection of anti-borrelial antibodies in the blood of sheep (e.g. Hovmark et al., 1986b; Fridriksdottir et al., 1992; Ciceroni et al., 1996), but, as discussed above, the same caveats regarding the specificity of the tests apply here. The susceptibility of sheep to systemic infection appears to be low (Stuen and Fridriksdottir, 1991), which limits the ability of B. burgdorferi s.l. to cause disease and also reduces the proportion of infected ticks in the immediate environment (Gray et al., 1995), thus limiting the risk to farmers. There is some evidence that adult ticks may become infected as the result of the transfer of the pathogens from infected to uninfected nymphal ticks co-feeding on sheep (Ogden et al., 1997), but the impact that this might have on the ecology and epidemiology of the disease is unknown.

Lyme borreliosis in other livestock Antibodies to B. burgdorferi s.l. have been detected in the blood of goats (Doby and Chevrier, 1990; Ciceroni et al., 1996) but there are no reports of LB in these animals. Seroprevalence in domestic pigs has not been investigated, though wild boar have been found to be seropositive (Doby et al., 1991). Studies on the susceptibility of deer to B. burgdorferi s.l. suggest that infected animals do not exhibit clinical signs though seroconversion occurs (Lane et al., 1994; Luttrell et al., 1994). Chickens have been infected experimentally and it was found that these

16

G. Stanek et al.

animals quickly became immune to B. burgdorferi s.s. and did not show any clinical symptoms (Piesman et al., 1996). More recent studies have shown that pheasants can function as reservoir hosts of B. garinii and Borrelia valaisiana in the UK (Kurtenbach et al., 1998), but no symptoms of disease in infected birds have been reported. In summary, there is no doubt that LB occurs in some domestic animals, and serological studies suggest that infection with B. burgdorferi s.l. is common. However, at present there is little evidence that LB is responsible for frequent and significant morbidity in any domestic animal species, with the possible exception of the dog. Diagnosis of LB in domestic animals is particularly problematic because of the lack of case definitions and of quality-assured laboratory tests.

Diagnostic Problems Clinical diagnosis of Lyme borreliosis Clinical signs and symptoms should always form a basis for diagnosis of human LB (Strle, 1999). Thus, a solid knowledge of clinical features is essential and case definitions (Stanek et al., 1996; Anon., 1997; Nadelman and Wormser, 1998) could be of help in everyday clinical practice and especially for comparison of the findings of different authors or groups. Caution should be used when clinical and laboratory experience from one side of the Atlantic Ocean is applied to patients on the other (Strle et al., 1999). A characteristic sign that enables diagnosis of LB in clinical practice is the typical EM skin lesion, and highly suggestive manifestations are ear lobe lymphocytoma, ACA and Bannwarth syndrome. The majority of the many other signs and symptoms are only suggestive and when expressed individually may have a very limited value for clinical diagnosis.

Laboratory evidence of borrelial infection Laboratory confirmation of borrelial infection is, as a rule, needed for all manifestations of LB with the exception of typical EM (Tugwell et al., 1997; Nadelman and Wormser, 1998; Strle et al., 1999). In clinical practice indirect laboratory methods are widely used to support the diagnosis of borrelial infection. Determination of borrelial IgM and IgG antibodies by immunofluorescence assays or ELISA has not been standardized, and the correlation of results from different laboratories and/or different commercial tests may be poor. Immunoblotting may solve some of the many dilemmas but could (especially in Europe) add additional questions to the field in which many uncertainties still exist (Robertson et al., 2000). A clinician should understand the sensitivity and specificity of these tests, and be aware of the predictive values for certain clinical manifestations. However,

History and Characteristics of Lyme Borreliosis

17

in general, there have not been clear-cut answers to such questions. In addition, even if a clinician relied on the results of serological tests, several important problems remain with their interpretation. There is no doubt that we need (but do not have) standardized and dependable tests to ascertain borrelial infection, with the ability to distinguish between recent and long-lasting infection, and also between active and inactive infection. T-cell proliferation assay results are also difficult to interpret, and apparently the specificity and sensitivity of T-lymphocyte recognition by B. burgdorferi s.l. antigens remain controversial (Strle et al., 1999). There have also been several problems with tests for direct detection of borrelial infection. The reliability of methods other than culture to detect spirochaetes in tissue specimens is open to question. Although B. burgdorferi s.l. grows relatively well under laboratory conditions, the spirochaete is not easily recovered from clinical specimens other than skin biopsy samples of EM. Culture results are as a rule available only several weeks after the material has been obtained (Wilske and Preac-Mursic, 1993), which is not very desirable for clinicians or for their patients. Moreover, the procedure itself is time-consuming and relatively expensive. However, for a new or unusual clinical manifestation to be proclaimed as a sign of LB, borrelial infection should be firmly ascertained – isolation of B. burgdorferi s.l. from the tissue (organ) involved is the best possible approach to such cases. Sometimes the question arises as to which specimen has the highest chance for successful culture. For example, for isolation of B. burgdorferi s.l. from EM, a skin biopsy on the periphery of the lesion is usually performed (Strle et al., 1996). However, in reviewing previous work on this topic, conclusive evidence was not found that the peripheral aspect of the lesion represents the optimal site for biopsy. Results of a prospective study in which the culture isolation rate of B. burgdorferi s.l. from the centre and the margin of typical EM skin lesions was compared (Jurca et al., 1998) did not reveal any significant difference in the isolation rate. However, in a continuation of the study, a significant difference was found in favour of central biopsy sites in patients with homogeneous EM lesions (Jurca et al., 1998). Detection of borrelial DNA by PCR has been acclaimed for its high sensitivity and specificity. However, the value of detection of borreliae by this method in human tissues and fluids has not been unequivocally established (Tugwell et al., 1997; Nadelman and Wormser, 1998). In a 2-year prospective study undertaken in Ljubljana, the detection rate of borreliae by culture and PCR was compared using skin specimens from patients with untreated or treated skin lesions of EM and ACA. Results for 758 typical EM specimens showed positive results in 36% for culture in modified Kelly-Pettenkofer medium, 25% for PCR and 24% for culture in BSKII medium. The differences are statistically significant. The overall positivity rate for all methods combined was 54%, but few specimens were positive by all three methods (only 6%). Results for typical ACA were similar (Picken et al., 1997). These findings suggest that the distribution of spirochaetes in skin biopsies is not homogeneous and that, although possessing the potential to provide a rapid diagnosis, PCR is not more sensitive than culture for direct detection of borreliae in skin (Picken et al., 1997).

18

G. Stanek et al.

Treatment Basic principles of treatment A well-established principle in medicine maintains that a reliable diagnosis forms a basis for rational treatment. However, in LB, with its lack of clinical specificity compounded by the lack of standardized serological tests and direct detection procedures, a reliable diagnosis is usually not an easily achievable goal. In general, treatment with antibiotics is efficacious in all stages of LB and for all clinical manifestations; however, it has been most effective early in the course of the illness (Steere, 2001). Choice of antibiotic depends upon many factors including the efficacy, pharmacokinetic profile, side effects, expected compliance and price. The most effective antibiotic, the optimal dosage and the most appropriate duration of treatment have not been precisely determined for any of the many clinical manifestations of the disease. However, those uncertainties should not and cannot be the reason for the extent of the disparities in treatment recommendations that have been evident in the last few years. Some physicians treat only patients who fulfil all rigorous criteria for established LB. Others treat all poorly defined ‘chronic’ LB patients with repeated or prolonged courses of antibiotics, or sometimes even with combinations of antibiotics that are ineffective in vitro. In the last few years a trend of prolongation of treatment of LB has been observed. The reasons for this have been predominantly based on disappointments with the results of treatment according to current recommendation, sometimes on the basis of reports of anecdotal experiences and not on the results of controlled studies. However, it is reasonable to expect that the recommendations for treatment will continue to change and there are indications that shorter treatment regimens rather than longer ones will prove to be the correct approach (Klempner et al., 2001). Current shortcomings in therapy should be a stimulus for well-designed studies to answer these unresolved questions. There are several potential reasons for antibiotic treatment failure in LB. One of the possible causes for treatment failure is persistence of borreliae in tissues. There are clinical and experimental data, predominantly in European literature, showing that, after presumed adequate (that is, recommended) treatment, borreliae can persist in tissue (Preac-Mursic et al., 1996; Straubinger et al., 1997); however, it is difficult to assess the magnitude of this problem. In some cases a possible explanation for treatment failure could be irreversible tissue damage caused during active borrelial infection, or inflammation in association with this infection, or the induction of autoimmune mechanisms; in the last two cases treatment with an antibiotic will be ineffective. An important and probably a common cause for treatment failure is misdiagnosis. It is quite possible that treating a patient with non-specific symptoms such as arthralgia and myalgia and with apparently detectable borrelial serum antibodies will fail, because this may not be treatment of LB but of patients with certain serological test results.

History and Characteristics of Lyme Borreliosis

19

No reliable data support the use of antibiotics in asymptomatic persons with positive borrelial serological test results. A decision to treat with an antibiotic should be based on the presence of at least suggestive clinical signs and symptoms. Since IgM and IgG antibodies to B. burgdorferi s.l. may persist for years after clinical recovery (Wilske et al., 2000), the role of serology in measuring response to treatment is dubious (Nadelman and Wormser, 1998). According to present knowledge, repeated courses of (prolonged) antibiotic treatment are not indicated. In general, the main indication for intravenous therapy (usually with ceftriaxone) should be based on proven or suspected involvement of the central nervous system; for the large majority of other manifestations, oral therapy is adequate (Nadelman and Wormser, 1998). All antibiotics have potential adverse effects and some patients have had serious or fatal complications during treatment of poorly substantiated LB.

Treatment of Cases Two recent reviews have addressed this topic in some detail (Strle et al., 1999; Wormser et al., 2000). In central Europe, various antibiotics and antimicrobial chemotherapeutics are used for the therapy of the different manifestations of LB. The recommended duration of treatment ranges from 2 to 4 weeks. However, the necessary duration of the treatment of LB could be shorter and we have to await results of further studies on this subject. Patients with solitary EM and borrelial lymphocytoma are treated with amoxicillin, azithromycin, doxycycline, cefuroxime-axetil or phenoxymethylpenicillin but only very exceptionally with erythromycin. The usual duration of treatment is 14 days. Nervous system involvement and severe Lyme carditis are ordinarily treated with ceftriaxone or penicillin G for 2–3 weeks and only exceptionally orally with doxycycline or amoxicillin. Ceftriaxone, penicillin G, doxycycline or amoxicillin are used for the therapy of ACA and arthritis. Parenteral therapy is used when there is concomitant or peripheral central nervous system involvement and may also be useful for treatment of EM in pregnancy (Maraspin et al., 1996) and in immunodeficient patients. Duration of treatment is usually 3 weeks for ACA, 2 weeks for arthritis and 4 weeks in the case of oral therapy for arthritis. In most cases, Lyme arthritis may be cured by antibiotic therapy, but about 10% of Lyme arthritis patients do not respond satisfactorily (Franz et al., 1999).

Prophylaxis of Lyme borreliosis Chemoprophylaxis of LB, that is, administration of antibiotics after a tick bite to prevent infection, is not standard practice for several reasons. Firstly, the infection rate after a tick bite is low, between 1 and 2%, and, secondly, the rate of preventing eventual LB is not predictable (Stanek and Kahl, 1999). However, in a prospective study with adults bitten by ticks in endemic areas of North America,

20

G. Stanek et al.

a single dose of 200 mg doxycycline was effective for the prevention of LB, provided the antimicrobial was given within 72 h after the tick bite (Nadelman et al., 2001). Immunoprophylaxis of LB with a recombinant OspA vaccine was introduced in the USA in the mid-1990s. This vaccine is designed to prevent infection by destroying B. burgdorferi s.s. in the midgut of the feeding tick. Other vaccines are being designed and studied that would be appropriate in preventing infections with the more heterogeneous agents of LB in Europe (see Hayes and Schrider, Chapter 11).

References Ackermann, R., Rehse-Küpper, B., Gollmer, E. and Schmidt, R. (1988) Chronic neurologic manifestations of erythema migrans borreliosis. Annals of the New York Academy of Sciences 539, 16–23. Afzelius, A. (1910) Verhandlungen der Dermatologischen Gesellschaft zu Stockholm, 28 Oct 1909. Archives of Dermatology and Syphilis 101, 404. Anon. (1995) Report of WHO Workshop on Lyme Borreliosis Diagnosis and Surveillance. WHO/CDS/VPH/ 95.14-1, Warsaw, Poland, 220 pp. Anon. (1997) Case definitions for infectious conditions under public health surveillance: Lyme disease (revised 9/96). Mortality and Morbidity Weekly Report 46 (suppl. RR-10), 20–21. Appel, M.J., Allan, S., Jacobson, R.H., Lauderdale, T.L., Chang, Y.F., Shin, S.J., Thomford, J.W., Todhunter, R.J. and Summers, B.A. (1993) Experimental Lyme disease in dogs produces arthritis and persistent infection. Journal of Infectious Diseases 167, 651–664. Asbrink, E. and Hovmark, A. (1985) Successful cultivation of spirochetes from skin lesions of patients with erythema chronicum migrans Afzelius and acrodermatitis chronica atrophicans. Acta Pathologica Microbiologica et Immunologica Scandinavica (B) 93, 161–163. Asbrink, E., Hovmark, A. and Olsson, I. (1986) Clinical manifestations of acro-

dermatitis chronica atrophicans in 50 Swedish patients. Zentralblatt für Bakteriologie und Hygiene A 263, 253–261. Axford, J.S. and Rees, D.H.E. (eds) (1994) Lyme Borreliosis. Life Sciences Vol. 260, Plenum Press, New York, 344 pp. Bäferstedt, B. (1943) Über Lymphadenosis benigna cutis. Acta Dermato-Venereologica (Stockholm) Suppl. 24, 1–202. Bannwarth, A. (1941) Chronische lymphozytäre Meningitis, entzündliche Polyneuritis und ‘Rheumatismus’. Archiv Psychiatrischer Nerven krankheiten 113, 284–376. Baranton, G., Marti Ras, N.N. and Postic, D. (1998) Molecular epidemiology of the aetiological agents of Lyme borreliosis. Wiener Klinische Wochenschrift 110, 850–855. Barbour, A.G. (1984) Isolation and cultivation of Lyme disease spirochetes. Yale Journal of Biology and Medicine 57, 521–525. Barbour, A.G. and Fish, D. (1993) The biological and social phenomenon of Lyme disease. Science 260, 1610–1616. Barbour, A.G. and Schrumpf, M.E. (1986) Polymorphisms of major surface proteins of Borrelia burgdorferi. Zentralblatt für Bakteriologie und Hygiene A 263, 83–91. Benach, J.L. and Bosler, E.M. (eds) (1988) Lyme disease and related disorders. Annals of the New York Academy of Sciences 539, 1–513. Benach, J.L., Bosler, E.M., Hanrahan, J.P., Coleman, J.L., Bast, T.F., Habicht,

History and Characteristics of Lyme Borreliosis

G.S., Cameron, D.J., Ziegler, J.L., Burgdorfer, W., Barbour, A.G., Edelman, R. and Kaslow, R.A. (1983) Spirochetes isolated from the blood of two patients with Lyme disease. New England Journal of Medicine 308, 740–742. Blowey, R.W., Carter, S.D., White, A.G. and Barnes, A. (1994) Borrelia burgdorferi infections in UK cattle: a possible association with digital dermatitis. Veterinary Record 135, 577–578. Borko, C. and Bole-Hribovsek, V. (1998) Infection rate of the tick Ixodes ricinus (Ixodidae, Acarina) with bacteria Borrelia burgdorferi sensu lato and in connection with the infection rate in grazing cattle. Zbornik Veterinarske Fakultete Univerza Ljubljana 36, 173–180. Breier, F.H., Aberer, E., Stanek, G., Khanakah, G., Schlick, A. and Tappeiner, G. (1999) Isolation of Borrelia afzelii from circumscribed scleroderma. British Journal of Dermatology 140, 925–930. Breier, F., Khanakah, G., Stanek, G., Kunz, G., Aberer, E., Schmidt, B. and Tappeiner, G. (2001) Isolation and polymerase chain reaction typing of Borrelia afzelii from a skin lesion in a seronegative patient with generalized ulcerating bullous lichen sclerosus et atrophicus. British Journal of Dermatology 144, 387–392. Browning, A., Carter, S.D., Barnes, A., May, C. and Bennett, D. (1993) Lameness associated with Borrelia burgdorferi infection in the horse. Veterinary Record 132, 610–611. Buchwald, A. (1883) Ein Fall von diffuser idiopathischer Haut Atrophie. Archives of Dermatology and Syphilis 10, 553–556. Burckhardt, J.L. (1911) Zur Frage der Follikel- und Keimzentrenbildung in der Haut. Frankfurter Zeitschrift für Pathologie 6, 352–359. Burgdorfer, W., Barbour, A.G., Hayes, S.F., Benach, J.L., Grunwaldt, E. and Davis, J.P. (1982) Lyme disease – a tick-

21

borne spirochetosis? Science 216, 1317–1319. Burgess, E.C. (1988) Borrelia burgdorferi infection in Wisconsin horses and cows. Annals of the New York Academy of Sciences 539, 235–243. Burgess, E.C. (1992) Experimentally induced infection of cats with Borrelia burgdorferi. American Journal of Veterinary Research 53, 1507–1511. Burgess, E.C. and Mattison, M. (1987) Encephalitis associated with Borrelia burgdorferi infection in a horse. Journal of the American Veterinary Medical Association 191, 1457–1458. Burgess, E., Gillette, D. and Pickett, J.P. (1986) Arthritis and panuveitis as manifestations of Borrelia burgdorferi infection in a Wisconsin pony. Journal of the American Veterinary Medical Association 186, 1340–1342. Burgess, E.C., Gendron-Fitzpatrick, A. and Wright, W.O. (1987) Arthritis and systemic disease caused by Borrelia burgdorferi infection in a cow. Journal of the American Veterinary Medical Association 191, 1468–1470. Cevenini, R., Sambri, V. and La Placa, M. (eds) (1994) Advances in Lyme Borreliosis Research. Societa Editrice Esculapio, Bologna, 239 pp. Chang, Y., Novosol, V., McDonough, S.P., Chang, C.F., Jacobson, R.H., Divers, T., Quimby, F.W., Shin, S. and Lein, D.H. (1999) Vaccination against Lyme disease with recombinant Borrelia burgdorferi outer-surface protein A (rOspA) in horses. Vaccine 18, 540–548. Chang, Y.E., Nosovol, V., McDonough, S.P., Chang, C.F., Jacobson, R.H., Divers, T., Quimby, F.W., Shin, S. and Lein, D.H. (2000) Experimental infection of ponies with Borrelia burgdorferi by exposure to ixodid ticks. Veterinary Pathology 37, 68–76. Choi, B.K., Natterman, H., Grund, S., Haider, W. and Gobel, U.B. (1997) Spirochetes from digital dermatitis lesions in cattle closely related to

22

G. Stanek et al.

treponemes associated with human periodontitis. International Journal of Systematic Bacteriology 47, 175–181. Christen, H.J., Hanefeld, F., Eiffert, H. and Thomssen, R. (1993) Epidemiology and clinical manifestations of Lyme borreliosis in childhood. Acta Paediatrica 82 (suppl. 386), 1–76. Ciceroni, L., Simeoni, J., Pacetti, A.I., Ciarrocchi, S. and Cacciapuoti, B. (1996) Antibodies to Borrelia burgdorferi in sheep and goats. Alto Adige-South Tyrol, Italy. Microbiologica 19, 171–174. Dambach, D.M., Smith, C.A., Lewis, R.M. and van Winkle, T.J. (1997) Morphologic, immunohistochemical, and ultrastructural characterization of a distinctive renal-lesion in dogs putatively associated with Borrelia burgdorferi infection: 49 cases (1987–1992). Veterinary Pathology 34, 85–96. Demirkan, I., Walker, R.I., Murray, R.D., Blowey, R.W. and Carter, S.D. (1999) Serological evidence of spirochaetal infections associated with digital dermatitis in dairy cattle. Veterinary Journal 157, 69–77. Doby, J.M. and Chevrier, S. (1990) Recherche des anticorps anti-Borrelia burgdorferi chez les caprins en Bretagne. Bilan serologique sur 602 chevres. Recueil de Médecine Vétérinaire 166, 799–804. Doby, J.M., Betremieux, C., Rolland, C. and Barrat, J. (1991) Les grands mammifères forestiers, réservoirs de germe pour Borrelia burgdorferi, agent de la maladie de Lyme? Recherche des anticorps chez 543 cervides et suides. Recueil de Médecine Vétérinaire 167, 55–61. Egenvall, A., Franzen, P., Gunnarsson, A., Engvall, E.O., Vagsholm, I., Wikstrom, U.B. and Artursson, K. (2001) Crosssectional study of the seroprevalence to Borrelia burgdorferi sensu lato and granulocytic Ehrlichia spp. and demographic, clinical and tick-exposure factors in Swedish horses. Preventive Veterinary Medicine 49, 191–208.

Font, A., Closa, J.M. and Mascort, J. (1992) Lyme disease in dogs in Spain. Veterinary Record 130, 227–228. Franz, J.K., Priem, S., Rittig, M.G., Burmester, G.R. and Krause, A. (1999) Studies on the pathogenesis and treatment of Lyme arthritis. Wiener Klinische Wochenschrift 111, 981–984. Fridriksdottir, V., Overnes, G. and Stuen, S. (1992) Suspected Lyme borreliosis in sheep. Veterinary Record 130, 323–324. Garin, C. and Bujadoux, R. (1922) Paralysie par les tiques. Journal de Médecine de Lyon 71, 765–767. Gerhards, H. and Wollanke, B. (1996) Antikorpertiter gegen Borrelien bei Pferden im Serum und im Auge und Vorkommen der equinen rezidivierenden Uveitis (ERU). Berliner und Münchener Tierarztliche Wochenschrift 109, 273–278. Gern, L., Estrada-Pena, A., Frandsen, F., Gray, J.S., Jaenson, T., Jongejan, F., Kahl, O., Korenberg, E., Mehl, R. and Nuttall, P. (1998) European reservoir hosts of Borrelia burgdorferi s.l. Zentralblatt für Bakteriologie 283, 196–204. Gibson, M.D, Omran, M.T. and Young, C.R. (1995) Experimental feline Lyme borreliosis as a model for testing Borrelia burgdorferi vaccines. Advances in Experimental Medicine and Biology 383, 73–82. Götz, H. (1954) Die Acrodermatitis chronica atrophicans Herxheimer als Infektionskrankheit. Hautarzt 5, 491–504. Götz, H. (1955) Die Acrodermatitis chronica atrophicans Herxheimer als Infektionskrankheit. Hautarzt 6, 294–252. Grann, D. (2001) Stalking Dr. Steere over Lyme disease. The New York Times, Magazine Desk, 17 June 2001. Gray, J.S. (1999) Tick-borne pathogen interactions. Infectious Disease Review 1, 117–121. Gray, J.S. and Stanek, G. (eds) (1998) Risk assessment in Lyme Borreliosis, a ‘concerted action’ in the EU Biomedicine and Health Programme (1993–1996). Zentralblatt für Bakteriologie und Hygiene 287, 175–269.

History and Characteristics of Lyme Borreliosis

Gray, J.S., Kahl, O, Janetzki, C, Stein, J. and Guy, E. (1995) The spatial distribution of Borrelia burgdorferi-infected Ixodes ricinus in the Connemara region of Co. Galway, Ireland. Experimental and Applied Acarology 19, 163–172. Hahn, C.N., Mayhew, I.G., Whitwell, K.E., Smith, K.C., Carey, D., Carter, S.D. and Read, R.A. (1996) A possible case of Lyme borreliosis in a horse in the UK. Equine Veterinary Journal 28, 84–88. Hauser, W. (1955) Zur Kenntnis der Akrodermatitis chronica atrophicans. Archives of Dermatology and Syphilis 199, 350–393. Hellerström, S. (1930) Erythema chronicum migrans Afzelii. Acta DermatoVenereologica (Stockholm) 11, 315–321. Herxheimer, K. and Hartmann, K. (1902) Über Acrodermatitis chronica atrophicans. Archives of Dermatology and Syphilis 61, 57–76. Herzer, P. (1991) Joint manifestations of Lyme borreliosis in Europe. Scandinavian Journal of Infectious Diseases 77, 55–63. Hollström, E. (1958) Penicillin treatment of erythema chronicum migrans Afzelius. Acta Dermato-Venereologica (Stockholm) 38, 285–289. Hopf, H.C. (1966) Acrodermatitis chronica atrophicans (Herxheimer) und Nervensystem. In: Monographien aus dem Gesamtgebiet der Neurologie und Psychiatrie, Vol. 114, Springer, Berlin, pp. 1–130. Hörstrup, P. and Ackermann, R. (1973) Durch Zecken übertragene Meningopolyneuritis (Garin-Bujadoux, Bannwarth). Fortschritte der Neurologie Psychiatrie 41, 583–606. Hovius, J.W., Hovius, K.E., Oei, A., Houwers, D.J. and van Dam, A.P. (2000) Antibodies against specific proteins of and immobilizing activity against three strains of Borrelia burgdorferi sensu lato can be found in symptomatic but not in infected asymptomatic dogs. Journal of Clinical Microbiology 38, 2611–2621.

23

Hovius, K.E. (1999) Borrelia infections in dogs: epidemiological, clinical and diagnostic aspects. PhD thesis, University of Utrecht, The Netherlands. Hovius, K.E., Stark, L.A., BleuminkPluym, N.M., van de Pol, I., Verbeekde Kruif, N., Rijpkema, S.G., Schouls, L.M. and Houwers, D.J. (1999) Presence and distribution of Borrelia burgdorferi sensu lato species in internal organs and skin of naturally infected symptomatic and asymptomatic dogs, as detected by polymerase chain reaction. Veterinary Quarterly 21, 54–58. Hovmark, A., Asbrink, E. and Olsson, I. (1986a) The spirochetal etiology of lymphadenosis benigna cutis solitaria Afzelius. Acta Dermato-Venereologica (Stockholm) 66, 479–484. Hovmark, A., Asbrink, E., Schwan, O., Hederstedt, B. and Christensson, D. (1986b) Antibodies to Borrelia spirochetes in sera from Swedish cattle and sheep. Acta Veterinaria Scandinavica 27, 479–485. Isogai, H., Isogai, E., Masuzawa, T., Yanagihara, Y., Matsubara, M., Shimanuki, M., Seta, T., Fukai, K., Kurosawa, N. and Enokidani, M. (1992) Seroepidemiological survey for antibody to Borrelia burgdorferi in cows. Microbiology and Immunology 36, 1029–1039. Johnson, R.C., Schmid, G.P., Hyde, F.W., Steigerwaldt, A.G. and Brenner, D.J. (1984) Borrelia burgdorferi sp.nov.: etiologic agent of Lyme disease. International Journal of Systematic Bacteriology 34, 496–497. Jurca, T., Ruzic-Sabljic, E., Lotric-Furlan, S., Maraspin, V., Cimperman, J. and Picken R.N. (1998) Comparison of peripheral and central biopsy sites for the isolation of Borrelia burgdorferi sensu lato from erythema migrans skin lesions. Clinical Infectious Disease 27, 636–638. Käsbohrer, A. and Schönberg, A. (1990) Serologische Untersuchungen zum Vorkommen von Borrelia burgdorferi bei

24

G. Stanek et al.

Haustieren in Berlin (West). Berliner und Munchener Tierarztliche Wochenschrift 103, 374–378. Klempner, M.S., Hu, L.T., Evans, J., Schmid, C.H., Johnson, G.M., Trevino, R.P., Norton, D., Levy, L., Wall, D., McCall, J., Kosinski, M. and Weinstein, A. (2001) Two controlled trials of antibiotic treatment in patients with persistent symptoms and a history of Lyme disease. New England Journal of Medicine 345, 85–92. Kornblatt, A.N., Urband, P.H. and Steere, A.C. (1985) Arthritis caused by Borrelia burgdorferi in dogs. Journal of the American Veterinary Medicine Association 186, 960–964. Kristoferitsch, W. (1989) Neuropathien bei Lyme-Borreliose. Springer Verlag, Vienna, 207 pp. Kristoferitsch, W., Sluga, E., Graf, M., Partsch, H., Neumann, R., Stanek, G. and Budka, H. (1988) Neuropathy associated with acrodermatitis chronica atrophicans: clinical and morphological features. Annals of the New York Academy of Sciences 539, 35–45. Kurtenbach, K., Peacey, M., Rijpkema, S.G.T., Hoodless, A.N., Randolph, S.E. and Nuttall, P.A. (1998) Differential transmission of genospecies of Borrelia burgdorferi sensu lato by game birds and small rodents in England. Applied and Environmental Microbiology 64, 1169–1174. Lane, R.S., Berger, D.M., Casher, L.E. and Burgdorfer, W. (1994) Experimental infection of Columbian black-tailed deer with the Lyme disease spirochete. Journal of Wildlife Diseases 30, 20–28. Levy, S.A. and Duray, P.H. (1988) Complete heart block in a dog seropositive for Borrelia burgdorferi. Similarity to human Lyme carditis. Journal of Veterinary International Medicine 2, 138–144. Liebisch, G. and Liebisch, A. (1999) Lymeborreliosis in dogs: risk of infection, interpretation of laboratory diagnosis

and vaccination. Praktische Tierarzt 80, 404. Liebisch, G., Assman, G. and Liebisch, A. (1999) Infection with Borrelia burgdorferi s.l. causes disease (Lyme borreliosis) in horses in Germany. Praktische Tierarzt 80, 498. Lipschütz, B. (1913) Über eine seltene Erythemform (Erythema chronicum migrans). Archives of Dermatology and Syphilis 118, 349–356. Lipschütz, B. (1923) Weiterer Beitrag zur Kenntnis der ‘Erythema chronicum migrans’. Archives of Dermatology and Syphilis 143, 365–374. Lischer, C.J., Leutenegger, C.M., Braun, U. and Lutz, H. (2000) Diagnosis of Lyme disease in two cows by the detection of Borrelia burgdorferi DNA. Veterinary Record 146, 497–499. Lissman, B.A., Bosler, E.M., Camay, H., Ormiston, B.G. and Benach, J.L. (1984) Spirochete-associated arthritis (Lyme disease) in a dog. Journal of the American Veterinary Medicine Association 185, 219–220. Logar, M., Lotric-Furlan, S., Maraspin, V., Cimperman, J., Jurca, T., RuzicSabljic, E. and Strle, F. (1999) Has the presence or absence of Borrelia burgdorferi sensu lato as detected by skin culture any influence on the course of erythema migrans? Wiener Klinische Wochenschrift 111, 945–950. Luttrell, M.P., Nakagaki, K., Howerth, E.W., Stallknecht, D.E. and Lee, K.A. (1994) Experimental infection of Borrelia burgdorferi in white-tailed deer. Journal of Wildlife Diseases 30, 146–154. Magnarelli, L.A., Anderson, J.F., Shaw, E., Post, J.E. and Palka, F.C. (1988) Borreliosis in northeastern United States. American Journal of Veterinary Research 49, 359–362. Magnarelli, L.A., Anderson, J.F., Levine, H.R. and Levy, S.A. (1990) Tick parasitism and antibodies to Borrelia burgdorferi in cats. Journal of the American Veterinary Medical Association 197, 63–66.

History and Characteristics of Lyme Borreliosis

Magnarelli, L.A., Ijdo, J.W., van Andel, A.E., Wu, C., Padula, S.J. and Fikrig, E. (2000) Serologic confirmation of Erhrlichia equi and Borrelia burgdorferi infections in horses from the northeastern United States. Journal of the American Veterinary Medical Association 217, 1045–1050. Manion, T.B., Khan, M.I., Dinger, J. and Bushmich, S.L. (1998) Viable Borrelia burgdorferi in the urine of two clinically normal horses. Journal of Veterinary Diagnostic Investigation 10, 196–199. Maraspin, V., Cimperman, J., LotricFurlan, S., Pleterski-Rigler, D. and Strle, F. (1996) Treatment of erythema migrans in pregnancy. Clinical Infectious Disease 22, 788–793. McKenna, P., Clement, J., van Dijck, D., Lauwerys, M., Carey, D., van den Bogaard, T. and Bigaignon, G. (1995) Canine Lyme disease in Belgium. Veterinary Record 136, 244–247. Müller, I., Khanakah, G., Kundi, M. and Stanek, G. (2002) Horses and Borrelia: immunoblot patterns with five Borrelia burgdorferi sensu lato strains and sera from horses of various stud farms in Austria and from the Spanish Riding School in Vienna. International Journal of Medical Microbiology 291, Suppl. 33 pp. 80–87. Nadelman, R.B. and Wormser, G.P. (1998) Lyme borreliosis. Lancet 352, 557–565. Nadelman, R.B., Nowakowski, J., Fish, D., Falco, R.C., Freeman, K., McKenna, D., Welch, P., Marcus, R., AgueroRosenfeld, M.E., Dennis, D.T. and Wormser, G.P. (2001) Prophylaxis with single-dose doxycycline for the prevention of Lyme disease after an Ixodes scapularis tick bite. New England Journal of Medicine 345, 79–84. Ogden, N.H., Nuttall, P.A. and Randolph, S.E. (1997) Natural Lyme disease cycles maintained via sheep by co-feeding ticks. Parasitology 115, 591–599. Oliver, J.H. Jr, Magnarelli, L.A., Hutcheson, H.J. and Anderson, J.F. (1999) Ticks and antibodies to Borrelia

25

burgdorferi from mammals at Cape Hatteras, NC and Assateague Island, MD and VA. Journal of Medical Entomology 36, 578–587. Overduin, L.M. and van den Bogaard, A.E. (1997) Lyme borreliosis in dogs. Tijdschrift voor Diergeneeskunde 122, 7–9. Parker, J.L. and White, K.K. (1992) Lyme borreliosis in cattle and horses: a review of the literature. Cornell Veterinarian 82, 253–274. Picken, M.M., Picken, R.N., Han, D., Cheng, Y., Ruzic-Sabljic, E., Cimperman, J., Maraspin, V., LotricFurlan, S. and Strle, F. (1997) A two year prospective study to compare culture and polymerase chain reaction amplification for the detection and diagnosis of Lyme borreliosis. Molecular Pathology 50, 186–193. Picken, R.N., Strle, F., Picken, M.M., Ruzic-Sabljic, E., Maraspin, V., LotricFurlan, S. and Cimperman, J. (1998) Identification of three species of Borrelia burgdorferi sensu lato (B. burgdorferi sensu stricto, B. garinii, and B. afzelii) among isolates from acrodermatitis chronica atrophicans lesions. Journal of Investigative Dermatology 110, 211–214. Piesman, J., Dolan, M.C., Schriefer, M.E. and Burkot, T.R. (1996) Ability of experimentally infected chickens to infect ticks with the Lyme disease spirochete, Borrelia burgdorferi. American Journal of Tropical Medicine and Hygiene 54, 294–298. Post, J.E., Shaw, E.E. and Wright, S.D. (1988) Suspected borreliosis in cattle. Annals of the New York Academy of Sciences 539, 488. Preac-Mursic, V., Wilske, B., Schierz, G., Pfister, H.W. and Einhäupl, K. (1984) Repeated isolation of spirochetes from the cerebrospinal fluid of a patient with meningoradiculitis Bannwarth. European Journal of Clinical Microbiology 3, 564–565. Preac-Mursic, V., Marget, W., Busch, U., Pleterski-Rigler, D. and Hagl, S. (1996) Kill kinetics of Borrelia burgdorferi and

26

G. Stanek et al.

bacterial findings in relation to treatment of Lyme borreliosis. Infection 24, 9–16. Priem, S., Burmester, G.R., Kamradt, T., Wolbart, K., Rittig, M.G. and Krause, A. (1998) Detection of Borrelia burgdorferi by polymerase chain reaction in synovial membrane, but not in synovial fluid from patients with persisting Lyme arthritis after antibiotic therapy. Annals of Rheumatic Diseases 57, 118–121. Robertson, J., Guy, E., Andrews, N., Wilske, B., Anda, P., Granstrom, M., Hauser, U., Moosmann, Y., Sambri, V., Schellekens, J., Stanek, G. and Gray, J. (2000) A European multicenter study of immunoblotting in serodiagnosis of Lyme borreliosis. Journal of Clinical Microbiology 38, 2097–2102. Rogers, A.B., Smith, R.D. and Kakoma, I. (1999) Serologic cross-reactivity of antibodies against Borrelia theileri, Borrelia burgdorferi, and Borrelia coriaceae in cattle. American Journal of Veterinary Research 60, 694–697. Rothwell, J.T., Christie, B.M., Williams, C. and Walker, K.H. (1989) Suspected Lyme disease in a cow. Australian Veterinary Journal 66, 296–298. Sälde, H. (1946) Erythema chronicum migrans Afzelius with meningitis. Acta Dermato-Venereologica (Stockholm) 31, 227–234. Schaltenbrand G. (1967) Durch Arthropoden übertragene Erkrankung der Haut und des Nervensystems. Verhandlungen der Deutschen Gesellschaft für Kreislaufforschung 72, 975–1006. Sköldenberg, B. and Stiernsted, G. (eds) (1991) Lyme Borreliosis 1990. Scandinavian Journal of Infectious Diseases, suppl. 77, 157 pp. Speck, S., Reiner, D. and Wittenbrink, M.M. (2001) Isolation of Borrelia afzelii in a dog. Veterinary Record 149, 19–20. Stanek, G. and Kahl, O. (1999) Chemoprophylaxis for Lyme borreliosis? Zentralblatt für Bakteriologie 289, 655–665.

Stanek, G., Wewalka, G., Groh, G. and Neumann, R. (1985) Isolation of spirochetes from the skin of patients with erythema chronicum migrans in Austria. Zentralblatt für Bakteriologie und Hygiene A 260, 88–90. Stanek, G., Flamm, H., Barbour, A.G. and Burgdorfer, W. (eds) (1986) Lyme Borreliosis. Gustav Fischer Verlag, Stuttgart, 501 pp. Stanek, G., Kristoferitsch, W., Pletschette, M., Barbour, A.G. and Flamm, H. (eds) (1989) Lyme Borreliosis II. Zentralblatt für Bakteriologie, suppl. 18, 365 pp. Stanek, G., Klein, J., Bittner, R. and Glogar, D. (1990) Isolation of Borrelia burgdorferi from the myocardium of a patient with longstanding cardiomyopathy. New England Journal of Medicine 322, 249–252. Stanek, G., O’Connell, S., Cimmino, M., Aberer, E., Kristoferitsch, W., Granstrom, M., Guy, E. and Gray, J. (1996) European Union Concerted Action on Risk Assessment in Lyme Borreliosis: clinical case definitions for Lyme borreliosis. Wiener Klinische Wochenschrift 108, 741–747. Steere, A.C. (2001) Lyme disease. New England Journal of Medicine 345, 115–125. Steere, A.C. and Malawista, S.E. (1979) Cases of Lyme disease in the United States: locations correlated with distribution of Ixodes dammini. Annals of Internal Medicine 91, 730–733. Steere, A.C., Malawista, S.E., Snydman, D.R., Shope, R.E., Andiman, W.A., Ross, M.R. and Steele, F.W. (1977a) Lyme arthritis: an epidemic of oligoarticular arthritis in children and adults in three Connecticut communities. Arthritis and Rheumatism 20, 7–17. Steere, A.C., Malawista, S.E., Hardin, J.A., Ruddy, S., Askenase, P.W. and Andiman, W.A. (1977b) Erythema chronicum migrans and Lyme arthritis: the enlarging clinical spectrum. Annals of Internal Medicine 86, 685–698. Steere, A.C., Malawista, S.E., Newman,

History and Characteristics of Lyme Borreliosis

J.H., Spieler, P.N. and Bartenhagen, N.H. (1980) Antibiotic therapy in Lyme disease. Annals of Internal Medicine 93, 1–8. Steere, A.C., Grodzicki, R.L., Kornblatt, A.N., Craft, J.E., Barbour, A.G., Burgdorfer, W., Schmid, G.P., Johnson, E. and Malawista, S.E. (1983) The spirochetal etiology of Lyme disease. New England Journal of Medicine 308, 733–740. Steere, A.C., Schoen, R.T. and Taylor, E. (1987) The clinical evolution of Lyme arthritis. Annals of Internal Medicine 107, 725–731. Straubinger, R.K., Summers, B.A., Chang, Y.F. and Appel, M.J. (1997) Persistence of Borrelia burgdorferi in experimentally infected dogs after antibiotic treatment. Journal of Clinical Microbiology 35, 111–116. Straubinger, R.K., Straubinger, A.F., Summers, B.A., Jacobson, R.H. and Erb, H.N. (1998) Clinical manifestations, pathogenesis, and effect of antibiotic treatment on Lyme borreliosis in dogs. Wiener Klinische Wochenschrift 110, 874–881. Strle, F. (1999) Principles of the diagnosis and treatment of Lyme borreliosis. Wiener Klinische Wochenschrift 111, 911–915. Strle, F., Pleterski-Rigler, D., Stanek, G., Pejovnik-Pustinek, A., Ruzic, E. and Cimperman, J. (1992) Solitary borrelial lymphocytoma: report of 36 cases. Infection 20, 201–206. Strle, F., Preac-Mursic, V., Cimperman, J., Ruzic, E., Maraspin, V. and Jereb, M. (1993) Azithromycin versus doxycycline for treatment of erythema migrans: clinical and microbiological findings. Infection 21, 83–88. Strle, F., Nelson, J.A., Ruzic-Sabljic, E., Cimperman, J., Maraspin, V., LotricFurlan, S., Cheng, Y., Picken, M.M., Trenholme, G.M. and Picken, R.N. (1996) European Lyme borreliosis: 231 culture-confirmed cases involving

27

patients with erythema migrans. Clinical Infectious Disease 23, 61–65. Strle, F., Nadelman, B.N., Cimperman, J., Nowakowski, J., Picken, R.N., Schwartz, I., Maraspin, V., AgueroRosenfeld, M.E., Varde, S., LotricFurlan, S. and Wormser, G.P. (1999) Comparison of culture-confirmed erythema migrans caused by Borrelia burgdorferi sensu stricto in New York State and by Borrelia afzelii in Slovenia. Annals of Internal Medicine 130, 32–36. Stuen, S. and Fridriksdottir, V. (1991) Experimental inoculation of sheep with Borrelia burgdorferi. Veterinary Record 129, 315. Süss, J. and Kahl, O. (eds) (1997) Tick-borne Encephalitis and Lyme Borreliosis. 4th International Potsdam Symposium on Tick-borne Diseases, Pabst Science Publishers, Lengerich, Germany, 285 pp. Süss, J. and Kahl, O. (eds) (1999) Tickborne encephalitis and Lyme borreliosis. 5th International Potsdam Symposium on Tick-borne Diseases. Zentralblatt für Bakteriologie 289, 491–769. Süss, J., Kahl, O. and Dautel (eds) (2002) Tick-borne encephalitis and Lyme borreliosis. 5th International Potsdam Symposium on Tick-borne Diseases. Zentralblatt für Bakteriologie. Thyresson, N. (1949) The penicillin treatment of acrodermatitis chronica atrophicans (Herxheimer). Acta DermatoVenereologica (Stockholm) 29, 572–621. Tugwell, P., Dennis, D.T., Weinstein, A., Wells, G., Shea, B., Nichol, G., Hayward, R., Lightfoot, R., Baker, P. and Steere, A.C. (1997) Laboratory evaluation in the diagnosis of Lyme disease. Annals of Internal Medicine 127, 1109–1123. Tuomi, J., Rantamaki, L.K. and Tanskanen, R. (1998) Experimental infection of cattle with several Borrelia burgdorferi sensu lato strains; immunological heterogeneity of strains as

28

G. Stanek et al.

revealed in serological tests. Veterinary Microbiology 60, 27–43. Weber, K. (1974) Erythema-chronicummigrans-Meningitis – eine bakterielle Infektions-krankheit? München Medizinische Wochenschrift 116, 1993–1998. Weber, K. and Neubert, U. (1986) Clinical features of early erythema migrans disease and related disorders. Zentralblatt für Bakteriologie und Hygiene A 263, 209–228. Weber, K., Schierz, G., Wilske, B. and Preac-Mursic, V. (1985) Das Lymphozytom – eine Borreliose. Zeitschrift für Hautkrankheiten 69, 1585–1598. Wells, S.J., Trent, A.M., Robinson, R.A., Knutson, K.S. and Bey, R.F. (1993) Association between clinical lameness and Borrelia burgdorferi antibody in dairy cows. American Journal of Veterinary Research 54, 398–405. Wilske, B. and Preac-Mursic, V. (1993) Microbiological diagnosis of Lyme Borreliosis. In: Weber, K., Burgdorfer, W. and Schierz, G. (eds) Aspects of Lyme Borreliosis. Springer, Berlin, pp. 267–299. Wilske, B., Preac-Mursic, V., Schierz, G. and von Busch, K. (1986) Immunochemical and immunological analysis of European Borrelia burgdorferi strains. Zentralblatt für Bakteriologie und Hygiene A 263, 92–102.

Wilske, B., Zöller, L., Brade, V., Eiffert, H., Göbel, U.B. and Stanek, G. (2000) Qualitätsstandards in der mikrobiologischen Diagnostik, MIQ 12/2000 Lyme-Borreliose. Urban & Fischer, Munich, Jena, 59 pp. Wokke, J.H.J., van Gijn, J., Elderson, A. and Stanek, G. (1987) Chronic forms of Borrelia burgdorferi infection of the nervous system. Neurology 37, 1031–1034. Wormser, G.P. (1999) Vaccination as a modality to prevent Lyme disease. A status report. Infectious Diseases Clinic of North America 13, 135–148. Wormser, G.P., Nadelman, R.B., Dattwyler, R.J., Dennis, D.T., Shapiro, E.D., Steere, A.C., Rush, T.J., Rahn, D.W., Coyle, P.K., Persing, D.H., Fish, D. and Luft, B.J. (2000) Practice guidelines for the treatment of Lyme disease. Clinical Infectious Disease 31 (suppl. 1), S1–S14. Yanagihara, Y. and Masuzawa, T. (eds) (1994) Present Status of Lyme Borreliosis and Biology of Lyme Borrelia. University of Shizuoka Press, 230 pp. Yanagihara, Y. and Masuzawa, T. (eds) (1997) Emerging and Re-emerging Diseases Transmitted by Arthropod Vectors and Rodents. University of Shizuoka Press, 409 pp.

Ecological Research on Borrelia burgdorferi sensu lato: Terminology and Some Methodological Pitfalls

2

Olaf Kahl,1a Lise Gern,2 Lars Eisen3 and Robert S. Lane3 1

Institute of Applied Zoology/Animal Ecology, Free University of Berlin, 12163 Berlin, Germany; 2Institut de Zoologie, University of Neuchâtel, Emile Argand 11, 2007 Neuchâtel, Switzerland; 3Department of Environmental Science, Policy and Management, University of California, Berkeley, CA 94720, USA

Introduction Among arthropod-borne human diseases in the cool-temperate regions of the northern hemisphere Lyme borreliosis (LB) is the most commonly reported. The causative agents, Borrelia afzelii, Borrelia burgdorferi sensu stricto (s.s.) and Borrelia garinii, are spirochaetes belonging to the B. burgdorferi species complex, or B. burgdorferi sensu lato (s.l.). Almost all known Borrelia spp. circulate between ticks (Arachnida, Acari, Ixodoidea) and a wide variety of vertebrates, which may include different orders of mammals (mainly insectivores, rodents, lagomorphs and ungulates), some bird species and a few reptiles. Tick-borne relapsing fever borreliae are usually transmitted by soft ticks (Argasidae, genus Ornithodoros), whereas hard ticks (Ixodidae) transmit B. burgdorferi s.l. (Burgdorfer and Hayes, 1989). The epidemiologically most important ticks or transmitting B. burgdorferi s.l. to humans are Ixodes persulcatus (northern mid Asia), Ixodes scapularis (eastern North America), Ixodes pacificus (western North America) and Ixodes ricinus (Europe and some adjacent areas) (see Eisen and Lane, Chapter 4). Each of the three components of the ecological system: the pathogens, pathogen-transmitting ticks (vectors) and pathogen-transmitting tick hosts (usually called reservoir hosts), and also their interrelationships, are dependent on external influence by various biotic and abiotic factors (Fig. 2.1). For example, vector ticks are dependent upon the availability of suitable host individuals for a

Present address: Blackwell Verlag, Kurfürstendamm 57, 10707 Berlin, Germany.

© 2002 CAB International. Lyme Borreliosis: Biology, Epidemiology and Control (eds J. Gray, O. Kahl, R.S. Lane and G. Stanek)

29

30

O. Kahl et al.

each life stage and also need high microclimatic relative humidities during the often extended off-host phases. Host-seeking in ticks is usually a seasonal activity and is strongly dependent on environmental factors. The existence of B. burgdorferi s.l. requires the occurrence of vector ticks and also vertebrate populations that must be able to transmit borreliae efficiently to feeding ticks. Although humans often become infected with B. burgdorferi s.l. and subsequently may fall ill, they are dead-end hosts and they are not involved in the natural circulation of the pathogen. The main ecological influence of humans on B. burgdorferi s.l. is indirect, by their impact on landscape ecology and on the distribution and abundance of some tick hosts, for example deer. It became clear soon after the discovery of B. burgdorferi (Burgdorfer et al., 1982) that this organism was an important cause of human disease. Although specific ecological research began immediately in the early 1980s both in North America and Europe and shortly thereafter in the former USSR and Japan, many key ecological and epidemiological questions still remain unanswered, at least regionally and locally. The main variables in the ecological puzzle of LB are: (i) the different vector species, most of them occurring in regions with strongly varying macroclimatic conditions and differing host spectra; and (ii) the existence of several closely related Borrelia spp., referred to as genospecies, that can only be differentiated with molecular methods. Research in tick-borne diseases seeks

Biotic and abiotic environmental factors

Pathogen Ticks

Tick hosts

Biotic and abiotic environmental factors Important proven relationship Proven relationship Presumed relationship

Fig. 2.1. The ecological system of Lyme borreliosis.

Ecological Research on B. burgdorferi

31

to determine what environmental and behavioural factors influence the risk of human infection and disease. Answering these and other questions is of both major ecological and epidemiological relevance. Although it is undoubtedly a principal task to identify those ticks and vertebrates that maintain the natural transmission cycle and those ticks capable of transmitting pathogens to humans, it is also essential to have a standardized and plausible set of terms for describing the different kinds of functional involvement of ticks and vertebrates in the ecology of LB and other tick-borne diseases. In this chapter, a terminology is suggested that emphasizes clarity, consistency and simplicity, and is largely based on terms introduced by Pavlovsky (1966). Our concept tries to be broad and logical by also paying attention to the minor and even antagonistic roles of certain ticks and vertebrates in the circulation of tickborne pathogens. Thus, the ecological effect rather than the mechanism involved is emphasized, a distinction that is crucial. This chapter also evaluates borrelia detection methods and outlines different experimental approaches for determining the functional involvement of ticks and their hosts. Some major pitfalls in ecological research and some of the more common errors and faulty interpretations occurring in the literature are identified. For example, not all naturally infected ticks and vertebrates are capable of transmitting a given agent, nor does the presence of specific antibodies against the agent in vertebrates confirm their infectiousness for feeding ticks.

Ecological Terminology A large variety of qualitative and quantitative terms and phrases have been used to describe the multiple functions and attributes of ticks and their hosts in the ecology (epizootiology) of tick-borne zoonoses (e.g. Pavlovsky, 1966; Balashov, 1972; Ginsberg, 1993; Nuttall and Labuda, 1994; Sonenshine and Mather, 1994; Randolph and Craine, 1995). However, there has been inconsistency in the use of these terms, and a standardized and logical terminology would increase the clarity of communication. The system of terms presented here is intended to draw attention to the present confused situation and, in particular, to provide the basis for future discussions on this topic. The terms tick, tick host and vertebrate hereafter usually refer to species but, depending on the context, they can also be used in relation to populations and individuals.

Qualitative terms (categories) To transmit a given pathogen and to be considered a vector, a tick species must: (i) feed on infectious vertebrates; (ii) be able to acquire the pathogen during the blood meal; (iii) maintain it through one or more life stages (transstadial passage); and (iv) pass it on to other hosts when feeding again. Otherwise it is a

32

O. Kahl et al.

non-vector tick. In some cases tick infection may be acquired from another tick, if a male infects a female during mating (Alekseev and Dublinina, 1996) or females transovarially infect their offspring (Burgdorfer and Varma, 1967). These four successive crucial steps in the transmission of pathogens by vectors are considered in a system of qualitative categories (Fig. 2.2). Ticks that feed on infectious hosts may be categorized as pathogen-exposed ticks. Most pathogen detection methods used in ticks identify carrier ticks, which harbour a particular pathogen without necessarily being infective for a host. These methods are unable to differentiate vector ticks from carrier ticks. Numerous species of ixodid ticks have been shown to harbour B. burgdorferi s.l. naturally (see Chapter 4). Detection of spirochaetes in unfed nymphal or adult ticks clearly indicates that borreliae were passed transstadially from the larval or nymphal stage. At least seven such carrier tick species are unable to transmit spirochaetes to the host (see Eisen and Lane, Chapter 4) and may therefore be considered non-vector ticks. Infected vector ticks, however, are able to transmit the given pathogen to a host during their blood meal and are involved in the agent’s natural maintenance cycle. A comparable system of qualitative categories is presented for vertebrates (Fig. 2.3). To be considered a reservoir host, the vertebrate species concerned must be a source of infection for ticks. It must also fulfil similar successive transmission criteria that define a vector: (i) it must be fed on by infected vector ticks, at least occasionally; (ii) it must take up a critical number of infectious agents during an infectious tick bite; (iii) it must allow the pathogen to multiply and to Ticks

Non-exposed ticks

Pathogen-exposed ticks

Non-carrier ticks

Carrier ticks

Non-vector ticks

Vector ticks

Fig. 2.2. Categories describing the functional involvement of ticks in tick-borne zoonoses (key ecological terms are in a double frame).

Ecological Research on B. burgdorferi

33

Tick hosts Tick maintenance hosts

Non-exposed hosts

Non-carrier hosts

Barrier hosts

Pathogen-exposed hosts

Carrier hosts

Non-reservoir hosts

Reservoir hosts

Fig. 2.3. Categories describing the functional involvement of tick hosts in tickborne zoonoses (key ecological terms are in a double frame).

survive for some time in at least certain parts of its body; and last but not least (iv) it must allow the pathogen to find its way into other feeding ticks. Each of these four successive steps is crucial to make a tick host an active link in the chain of transmission and will therefore be considered in the following system of qualitative categories (Fig. 2.3). Tick hosts include all vertebrates that ticks feed on in nature. Local and regional tick abundance (or even tick occurrence) largely depends on the presence of one or more tick maintenance hosts (sometimes called ‘amplifying hosts’, a term also confusingly applied occasionally to vertebrates that transmit pathogens). The availability of suitable hosts (medium-sized and large mammals) for adults may be a limiting factor or bottle-neck for Ixodes spp. in some areas, for example in city parks or some suburban environments with an impoverished vertebrate fauna (Dautel and Kahl, 1999). Empirical studies have found correlations between local vector tick density and local abundance of deer both in Europe (Gray et al., 1992) and North America (reviewed by Wilson, 1998). Pathogen-exposed hosts are fed on by infected vector ticks, at least occasionally. Most pathogen detection methods used in tick hosts determine the functional category carrier hosts, i.e. hosts that harbour a particular pathogen for at least a short period but are not necessarily infective for ticks. These same methods do not distinguish carrier hosts from reservoir hosts, and at least in the case of B. burgdorferi s.l. the unwarranted assumption is

34

O. Kahl et al.

occasionally made that these categories are synonymous. ‘Susceptible hosts’ harbour, at least temporarily, live agents in a density suggesting that they could be infectious for feeding ticks but with experimental proof lacking. Reservoir hosts are proven natural hosts of vector ticks, and ticks may become infected while feeding on them. It is therefore common to all reservoir hosts that they increase the number of infected ticks in a particular area and thereby exert a positive ecological effect on the pathogen concerned. Infectivity lasts either lifelong, as in many small mammalian species when infected with B. afzelii or B. burgdorferi s.s., or for only some shorter period, as in certain small mammalian species infected with tick-borne encephalitis (TBE) virus for only a few days (Randolph et al., 1996). Uninfected ticks can also acquire an infection when feeding together with infected vector ticks on a ‘co-feeding transmission host’. Although the vertebrate is used only as a short-term vehicle by the agent, without showing an overt infection, co-feeding transmission may be ecologically very important for some tickborne pathogens, as postulated for TBE virus by Randolph et al. (1999). Because of the short-term infection after an infectious tick bite, a co-feeding transmission host’s significance depends strongly on the frequency with which infective and non-infected vector ticks co-feed on it (Randolph et al., 2000). Although the pathogen/host relationships in co-feeding transmission strongly differ from those in ‘classical’ reservoir host transmission, the same ecological effect of increasing the number of infected ticks in an area may justify the inclusion of ‘co-feeding transmission hosts’ under the well-established term ‘reservoir hosts’. Note that the term co-feeding transmission refers essentially to a transmission mechanism that may or may not be ecologically significant. Non-reservoir hosts may have contact with infected ticks and may or may not develop a long-lasting infection but they are incapable of transmitting the infection to ticks. Consequently, non-reservoir hosts do not directly support the circulation of the agent. Nevertheless, infected ticks may retain their infection when feeding on such a host and disseminate it later to another host or other ticks. As an example, white-tailed deer, Odocoileus virginianus, although apparently a non-reservoir host of B. burgdorferi s.s. (Telford et al., 1988), seem to be strongly implicated in the recent spread of that agent in parts of eastern North America, probably because of their function as maintenance hosts of the vector, I. scapularis (Spielman et al., 1993). Barrier hosts decrease the number of infected ticks, sometimes only after multiple contacts with vector ticks or the agent. Barrier host is a collective term describing an ecological effect rather than a particular mechanism. The mechanism by which the negative effect on the agent’s perpetuation is exerted can vary, for example, refractoriness to infection and elimination of infection in feeding ticks (zooprophylactic host), or destruction of large numbers of vector ticks (during or after feeding) because of, for example, efficient grooming or an acquired anti-tick immunity. Thus, it is important to know about tick performance (survival, developmental success, fecundity) after feeding on a given host species to assess its role in LB ecology. Confirmed barrier hosts for B. burgdorferi

Ecological Research on B. burgdorferi

35

s.s. include two North American lizard species, Sceloporus occidentalis and Elgaria multicarinata (Lane and Quistad, 1998; Wright et al., 1998). Qualitative terms not included in Figs 2.2 and 2.3 include bridging vectors, bridging hosts and experimental vectors/reservoir hosts. Bridging vectors transmit the agent to humans without necessarily being important vectors for the maintenance of the agent in nature (example: I. pacificus for B. burgdorferi s.s. in western North America, the non-anthropophilic Ixodes spinipalpis being the vector maintaining the enzootic cycle in small mammals). The term ‘bridging vector’ may relate only to a certain life stage, the one actually transmitting an infection to humans (example: female I. persulcatus for the TBE virus, for B. afzelii and B. garinii). Bridging hosts are naturally infested by more than one vector tick species in nature and may therefore form a link between distinct enzootic cycles of infection. Examples include the European hedgehog (Erinaceus europaeus) fed upon by I. ricinus and Ixodes hexagonus (Gern et al., 1997). Bridging hosts may be the source of infection for some bridging vectors, for example, for duskyfooted woodrats, Neotoma fuscipes, in California, which can be parasitized by the human-biting I. pacificus and the non-human-biting I. spinipalpis in the same area (Lane and Keirans, 1999). An experimental vector (tick) must have been shown to be capable of acquiring a particular agent while feeding, maintaining it until the following blood meal and subsequently transmitting it to a host via the next life stage, but is not necessarily involved in its natural circulation (e.g. the hard tick Rhipicephalus appendiculatus for the TBE virus (Labuda et al., 1993a,b)). If a tick species carrying B. burgdorferi s.l. in the field is shown to transmit borreliae in a feeding experiment there is strong evidence that it is a vector tick. Similarly, experimental reservoir hosts are capable of passing on the agent to feeding ticks in the laboratory but are not necessarily involved in its natural circulation, for example, the Mongolian gerbil, Meriones unguiculatus, for B. afzelii and B. burgdorferi s.s. (Kahl et al., 1998a; O. Kahl et al., unpublished data). If a vertebrate species carrying B. burgdorferi s.l. in the field is shown to transmit borreliae in a tick feeding experiment it is very likely to be a reservoir host. In both cases, experiments have to show whether pathogen transmission is possible for only one or a few highly adapted genotypes of a particular genospecies or whether this is also a common phenomenon when feral borreliae are used. The meanings of the terms ‘vector/reservoir competence’ have been diverse and often misleading in the literature and it might be extremely difficult to introduce new standardized definitions for such popular terms. We suggest that these terms should be used only qualitatively, describing the proven function of a particular species or population. Vector and reservoir competence are definitely not synonymous merely with the experimentally proven capability of transmission of a particular agent during tick feeding. The situation in nature must also be taken into account. The adjective ‘competent’ is superfluous, and instead of describing vectors or reservoirs as having high or low competence, we suggest describing them to be of high or low ‘importance’ on the basis of their general significance in the particular ecosystem concerned. This importance is a function

36

O. Kahl et al.

of their abundance in the habitat, their infectivity and their ‘capacity’ as described under ‘quantitative terms’ below. It is appropriate to distinguish primary versus secondary vectors and reservoirs, the former being ecologically important, the latter not. Ticks or vertebrates incapable of transmitting a pathogen are by definition not vectors or reservoirs of that pathogen. ‘Incompetent vector’ and ‘incompetent reservoir’ are contradictions in terms.

Quantitative terms There is even more confusion about the use of quantitative than of qualitative terms. Quantitative terms refer to the epizootiological and epidemiological consequences that arise from certain physiological capabilities as well as ecological and behavioural traits of ticks or their hosts in connection with the natural circulation of infectious agents. Quantitative terms relating to questing ticks and ticks on hosts have been reviewed by Wilson (1994). Briefly, abundance of questing ticks is a dimensionless relative measure, whereas density is an absolute measure describing the number of ticks per area unit. With regard to ticks on hosts, prevalence of infestation describes the percentage of examined hosts that are infested. Mean density of ticks per host is calculated as the total number of ticks observed divided by the total number of hosts examined, whereas intensity of infestation is the total number of ticks observed divided by the number of infested hosts. However, the distribution of ticks is typically highly aggregated within host populations, with most hosts harbouring a few ticks and a few hosts harbouring high numbers. Therefore the median value of infesting ticks may better describe the typical infestation level for an individual host. Finally, prevalence of infection refers to the percentage of examined ticks infected with a pathogen and intensity of infection describes the average number of pathogens per infected tick. Tick infectivity denotes the efficiency with which the infection is transmitted from a tick population or species to hosts. It can vary between different life stages and populations and also in relation to different vertebrate hosts. Specific tick infectivity denotes the efficiency with which the infection is transmitted from a tick population or species to a particular host species. The specific tick infectivity also may differ depending on the tick stage and phenotype of the agent involved (Dolan et al., 1998). Host infectivity denotes the efficiency with which the infection is transmitted from a tick host population or species to feeding ticks. The infectivity of a vertebrate may vary during its lifetime and between individuals due to physiological factors associated with ageing or immunological phenomena such as prior experience with ticks and tick-borne agents. Specific host infectivity denotes the proportion of ticks of a particular species and life stage that becomes infected during feeding. Also, the specific host infectivity may differ depending on the phenotype of the agent involved. Vector capacity describes the absolute contribution made by a particular

Ecological Research on B. burgdorferi

37

vector species to the natural prevalence of infection with a given pathogen in vertebrates in a certain area. Determination of the vector capacity of a tick should include quantitative data on the proportion of reservoir hosts naturally infested by the tick, the percentage of infected ticks, and the likelihood that one infected tick transmits the pathogen to a reservoir host (Randolph and Craine, 1995). Accordingly, relative vector capacity describes the proportional contribution made by a particular vector to the natural prevalence of infection in vertebrates in a certain area in relation to that of other possible vectors. Reservoir capacity describes the absolute contribution made by a particular reservoir or co-feeding transmission host to the natural prevalence of infection with a given pathogen in vector ticks in a certain area. Determination of the reservoir capacity of a vertebrate should include quantitative data on its density, its infectivity for feeding ticks, the duration of the infective period as well as vector tick infestation, vector tick feeding success and the subsequent tick moulting success (Mather et al., 1989; Manelli et al., 1993; Randolph and Craine, 1995). Accordingly, relative reservoir capacity (synonymous with reservoir potential, as used by Mather et al. (1989)) describes the proportional contribution made by a particular reservoir or co-feeding transmission host to the natural prevalence of infection with a given pathogen in vector ticks in a certain area in relation to that of other possible reservoir hosts. Empirical assessment of the relative reservoir capacity of various vertebrates in an area is an arduous task that few researchers have attempted, with few species considered (e.g. Mather et al., 1989; Tälleklint and Jaenson, 1994; Kurtenbach et al., 1995). Vector potential is a subjective assessment of the putative vector capacity of a given tick based on the results of incomplete laboratory experiments or insufficient field observations on, for example, its abundance, its tendency to feed on reservoir hosts, its feeding performance on reservoir hosts including its subsequent developmental success and/or its infectivity for different host taxa. Vector potential has some similarity with the collective term ‘vectorial capacity’ as used by Spielman et al. (1984), including all those vector-related variables affecting stability of pathogen transmission. Reservoir potential is a subjective assessment of the putative reservoir capacity of a given vertebrate based on the limited results of incomplete laboratory experiments or insufficient field observations on, for example, its abundance, its levels of infestation by different vector tick stages, its infectivity for vector ticks, and/or the duration of the infective period.

Pathogen Detection: Advantages and Limitations of Different Methods Care must be taken in choosing an adequate experimental approach to demonstrate that a particular tick or vertebrate belongs to a certain functional category in tick-borne zoonoses, or to answer a specific key ecological epidemiology question. In addition to the isolation of an agent from field-collected ticks and evidence of blood-feeding contact between suspected vectors and reservoir hosts,

38

O. Kahl et al.

in vivo transmission experiments are crucial for proving the capability of a vector or of a reservoir host (DeFoliart et al., 1987).

Identification of borrelia carriers The first step in an ecological epidemiology investigation is usually to identify one or more carriers of a given pathogen in a certain area. The next step typically is to assess the prevalence of infection in different parts of a tick population (i.e. different life stages, host-seeking or feeding individuals) over time. In addition to a proper spatial and temporal sampling pattern, reliable identification of the causative agents in ticks and their hosts is an important prerequisite. The detection method or combination of methods used must be sufficiently sensitive and specific. Several methods that can be used to identify carriers of B. burgdorferi s.l. offer different combinations of advantages and limitations and are summarized in Table 2.1.

Table 2.1. The strengths and limitations of Borrelia burgdorferi s.l. detection methods. Target

Detection method

Remarksa

Living (motile) cells

DFM PCM Culture

Detection of live spirochaetes, low specificity

Dead cells

IFA ELISA FISH

Detection of marked surface structures (IFA, ELISA) or genetic material (FISH), quantification of infection, high specificity possible

Genetic material

FISH PCR Sequencing

Detection of genetic material, high specificity

Specific antibodies in the vertebrate host

ELISA IFA Immunoblot

Determination of exposure, low to moderate specificity

a

The various methods may differ in their sensitivity, with culture, DFM and PCM probably being less sensitive, and IFA, PCR and FISH being more sensitive. However, there is no single method that is intrinsically more sensitive than the others. DFM, dark-field microscopy; PCM, phase-contrast microscopy; IFA, immunofluorescence assay; ELISA, enzyme-linked immunosorbent assay; FISH, fluorescence in situ hybridization; PCR, polymerase chain reaction.

Ecological Research on B. burgdorferi

39

Ticks There are three detection methods that demonstrate the viability and motility of borreliae: dark-field microscopy (DFM), phase-contrast microscopy (PCM) and culture. All three methods have a low specificity, so that it is not possible to differentiate between Borrelia species. Culture is too costly to be used for routine detection of borreliae but it affords a large number of cells that facilitates subsequent identification and characterization by PCR, monoclonal antibodies with differential diagnostic value and other sophisticated methods. However, culture is probably not possible with every B. burgdorferi s.l. isolate and may also result in a reduced genetic diversity after an increasing number of passages, due to selection for specific culture-adaptive genotypes (Norris et al., 1997). To save expense and time, DFM and culture can be combined so that only those ticks having a visible spirochaetal infection are cultivated. The remaining detection methods (see Table 2.1) detect only certain parts of the borrelial cell such as outer surface proteins, other proteins or genetic material. In doing so, they demonstrate infection but, strictly speaking, not the presence of living agents, although the detection of RNA by PCR may indicate living organisms since RNA degrades very rapidly. The immunofluorescence assay (IFA) may also indicate living borreliae by showing the typical helical morphology of the spirochaete. It is also possible to combine culture and IFA to obtain clear evidence of living borreliae. IFA is a simple, relatively inexpensive technique that can be as sensitive as PCR for detecting B. burgdorferi s.l. (e.g. Kahl et al., 1998b) and provides data on the intensity of infection, although some skill is required to eliminate subjectivity. If large numbers of ticks have to be examined quickly, IFA and PCR are the methods of choice because the samples can be fixed before investigation. However, differentiation of B. burgdorferi s.l. genospecies with IFA in a large number of ticks is, although possible, not practical. Ticks and their hosts, especially in Europe and Asia, may be co-infected with multiple genospecies of B. burgdorferi s.l. In central Europe, for example, I. ricinus populations harbour B. afzelii, B. garinii, B. burgdorferi s.s. and Borrelia valaisiana to varying degrees, and identification of mixed infections in ticks is possible with PCR (e.g. Rijpkema et al., 1995). Pooling groups of ticks, although cost-saving and offering some advantages from a biometric view (Abel et al., 1999), makes it impossible to determine the proportion of mixed infections. Because the tick moult seems to be a critical physiological challenge for borreliae in ticks (Piesman et al., 1990), unfed ticks found to be infected with borreliae under natural conditions are promising candidates as vectors. Partially or fully fed ticks removed from hosts that contain borreliae may or may not be vectors because nearly all haematophagous arthropods feeding on reservoir hosts of B. burgdorferi s.l. are likely to ingest some spirochaetes with the blood meal. Detection of infection in a particular arthropod or vertebrate is not indicative of infectiousness. This can only be proved by transmission experiments, as mentioned above.

40

O. Kahl et al.

Vertebrates Detection of infection in vertebrates can be based on host seroconversion, detection of spirochaetal DNA in host fluids or tissue, or recovery of live borreliae from cultured host tissue, e.g. ear-punch biopsies (Sinsky and Piesman, 1989). Host seroconversion demonstrates that the vertebrate was exposed to spirochaetes but does not confirm that a viable infection was established, nor does seropositivity indicate infectiousness, but it may be a first step towards describing the presence of a microorganism in an area. Likewise, detection of spirochaetal DNA in host tissue or fluids, although allowing detection of the genospecies involved, is not necessarily indicative of a viable infection. As previously suggested, detection of spirochaetal RNA might indicate living organisms, but this approach does not seem to have been adopted for detection of Borrelia spp. Xenodiagnosis offers an alternative to culture for certain hosts (Donahue et al., 1987) and is the method of choice for proving host infectivity to feeding ticks. The result is positive if spirochaetes are detected in previously spirochaete-free ticks after they have fed on either naturally or experimentally exposed hosts.

Identification of borrelia vectors and reservoir hosts None of the static tests capable of identifying carriers of borreliae is able to predict with certainty whether a given tick or vertebrate can transmit the agent, i.e. act as a vector or reservoir host. This can only be determined by carrying out additional transmission experiments with a well-defined pathogen. As noted by Lane (1994) and Randolph and Nuttall (1994), there are several difficulties in conducting biologically meaningful transmission studies. The variation sometimes observed in specific tick or host infectivity for different isolates of the same B. burgdorferi genospecies indicates that the use of more than one isolate in transmission studies would be preferable. Furthermore, because of the enlarging spectrum of B. burgdorferi genospecies (see Bergström et al., Chapter 3), it is especially difficult to draw the right conclusions from transmission experiments with a negative result. A vertebrate may be a reservoir host for one B. burgdorferi genospecies but not for another, as the example of different types of European and Asian B. garinii impressively demonstrates even within a particular genospecies (see Miyamoto and Masuzawa, Chapter 8). As one would expect, studies based on a single isolate that performs well in an experimental model may not always be representative of the specific tick or host infectivity for the whole genospecies. Problems related to the use of long-term cultured spirochaetes in studies (Schwan et al., 1988) can be diminished or avoided when using borreliae from a quasi-natural maintenance system in an experimental reservoir host species and a vector tick species in the laboratory. During feeding, nymphal vector ticks infect laboratory hosts, which subsequently transmit the infection to larval ticks.

Ecological Research on B. burgdorferi

41

A new transmission cycle (or vertebrate passage) begins with the feeding of the resultant hungry nymphs. Although it is possible to establish such a system by introducing an isolate from culture into a tick or a vertebrate, the feeding of field-collected vector ticks on a naive experimental reservoir host may form an interesting alternative, especially in areas where only one genospecies exists or where certain hosts only carry a single genospecies. Using this approach, a number of feral B. afzelii strains have been maintained for transmission experiments and one strain has completed its sixth gerbil passage after 5 years in the laboratory using I. ricinus as the vector (O. Kahl et al., unpublished data). These kinds of experimental studies can also begin with naturally infected feral hosts and xenodiagnostic vector ticks. Vectors The capability of a particular tick species to transmit a given agent must be shown by: (i) infecting uninfected ticks – usually larvae – during feeding on an infective host; (ii) allowing them to moult to the next life stage; (iii) feeding the resultant ticks – usually nymphs – on naive hosts; and (iv) demonstrating infection in the hosts after tick-feeding. Infecting ticks by artificial feeding (e.g. on a capillary tube containing living borreliae in culture medium) followed by feeding the ticks on a host is an unnatural situation. Agent transmission by in vitro-infected ticks must be confirmed by an in vivo approach. Transstadial survival is a critical challenge for a tick-borne agent and passing through all the different developmental phases (engorged phase, moult, unfed phase and feeding again) of the tick may allow the agent to synchronize its own physiology with that of the vector tick (Schwan et al., 1995; de Silva et al., 1996), and therefore should be part of any basic transmission experiment. Although survival of a pathogen from one life stage to another is commonly referred to as transstadial transmission there is in fact no transmission involved and a more accurate term would be transstadial passage since the tick simply stays infected through the moult. Reservoir hosts Xenodiagnosis is the classical method of choice for determining the host infectivity or specific host infectivity of a particular vertebrate species for a given agent. Feral animals captured in the field and caged for collection of detached ticks can be used and such an in vivo approach appears to be superior to the use of needle-infected animals (Gern et al., 1993; Piesman, 1993; Kurtenbach et al., 1994). Xenodiagnosis on feral animals may be problematic if infectivity in reservoir hosts and especially in co-feeding transmission hosts is short-lived, as in small mammals infected with TBE virus. It is very difficult in most geographical areas to determine the whole spectrum of reservoir hosts or the relative reservoir capacity of one particular or several vertebrate species. Blood-meal identification promises to be an elegant and effective tool for the identification of important reservoir hosts of B. burgdorferi s.l. and

42

O. Kahl et al.

other tick-borne agents (Kirstein and Gray, 1996). Briefly, each field-collected unfed nymphal vector tick is examined both for its B. burgdorferi s.l. infection status to genospecies level and for the source of its larval blood meal in order to identify the larval host. The latter can be done by amplifying PCR products of the cytochrome b gene from the remnants of the larval blood meal. It is thus possible to identify the vertebrate host source of the pathogen detected in the nymphal tick. This approach also takes account of the developmental success of vector ticks after they have fed on different host species, a factor whose importance can easily be underestimated in laboratory studies. By using this method, a list of candidate reservoir hosts can be identified with minimal trapping and handling of feral animals.

Concluding Remarks As is common for other tick-borne pathogens, the causative agents of LB, B. burgdorferi s.l., circulate between species of ticks and several tick hosts. A number of different terms exist for the description of the various functions and properties of ticks and their hosts in tick-borne zoonoses, but there has been inconsistency in their usage. A simple scheme has been presented here in an attempt to establish a logical set of well-defined ecological terms describing the various roles of ticks and tick hosts in LB and other tick-borne zoonoses, using previously established terms wherever possible. Vector or non-vector, reservoir host or nonreservoir host are the key qualitative questions in relation to the role played by ticks and their hosts, respectively, in the circulation of a particular zoonotic agent in nature. A further set of terms has been suggested that might help quantify the contribution of a given species or population to the natural circulation of a given zoonotic agent and to its ecological epidemiology importance. On the basis of all these definitions, the advantages and limitations of different experimental approaches for determining what particular functional category a tick or tick host belongs to have been outlined. Serodiagnosis in vertebrates only determines exposure to a pathogen. Although carrier ticks and carrier hosts may not be key functional categories, the great majority of detection methods identify just these categories but not infectiousness of the carrier. Only transmission experiments can help to clarify whether or not a given tick is a vector or non-vector and a given tick host is a reservoir host or non-reservoir host. It must be emphasized that the finding in a laboratory experiment that a particular tick or vertebrate species transmits a given agent does not mean that it is a vector or reservoir host under natural conditions. The species under investigation must be, at least sporadically, in contact with the agent in nature, usually by feeding on reservoir hosts in the case of vector ticks or by being fed on by vector ticks in the case of reservoir hosts.

Ecological Research on B. burgdorferi

43

Acknowledgement We are especially grateful to Jeremy Gray (Dublin, Irish Republic) for his contribution to the development and writing of this chapter.

References Abel, U., Schosser, R. and Süss, J. (1999) Estimating the prevalence of infectious agents using pooled samples: biometrical considerations. Zentralblatt für Bakteriologie 289, 550–563. Alekseev, A.N. and Dubinina, H.V. (1996) Exchange of Borrelia burgdorferi between Ixodes persulcatus (Acari: Ixodidae) sexual partners. Journal of Medical Entomology 33, 351–354. Balashov, Y.S. (1972) Bloodsucking ticks (Acari, Ixodidae) – vectors of diseases of man and animals. Miscellaneous Publications of the Entomological Society of America 8, 163–376. Burgdorfer, W. and Hayes, S.F. (1989) Vector–spirochete relationships in louse-borne and tick-borne borreliosis with emphasis on Lyme disease. In: Harris, K.F. (ed.) Advances in Disease Vector Research. Springer-Verlag, New York, pp. 127–150. Burgdorfer, W. and Varma, M.G.R. (1967) Trans-stadial and transovarial development of disease agents in arthropods. Annual Review of Entomology 12, 347–376. Burgdorfer, W., Barbour, A.G., Hayes, S.F., Benach, J.L., Grunwaldt, E. and Davis, J.P. (1982) Lyme disease – a tick-borne spirochetosis? Science 216, 1317–1319. Dautel, H. and Kahl, O. (1999) Ticks (Acari: Ixodoidea) and their medical importance in the urban environment. In: Robinson, W.H., Rettich, F. and Rambo, G.W. (eds) Proceedings of the 3rd International Conference on Urban Pests, Vol. 3. Grafické závody Hronov, Czech Republic, pp. 73–82. DeFoliart, G.R., Grimstad, P.R. and

Watts, D.M. (1987) Advances in mosquito-borne arbovirus/vector research. Annual Review of Entomology 32, 479–505. Dolan, M.C., Piesman, J., Mbow, M.L., Maupin, G.O., Péter, O., Brossard, M. and Golde, W.T. (1998) Vector competence of Ixodes scapularis and Ixodes ricinus (Acari, Ixodidae) for three genospecies of Borrelia burgdorferi. Journal of Medical Entomology 35, 465–470. Donahue, J.G., Piesman, J. and Spielman, A. (1987) Reservoir competence of white-footed mice for Lyme disease spirochetes. American Journal of Tropical Medicine and Hygiene 36, 92–96. Gern, L., Schaible, U.E. and Simon, M.M. (1993) Mode of inoculation of the Lyme disease agent Borrelia burgdorferi influences infection and immune responses in inbred strains of mice. Journal of Infectious Diseases 167, 971–975. Gern, L., Rouvinez, E., Toutoungi, L.N. and Godfroid, E. (1997) Transmission cycles of Borrelia burgdorferi sensu lato involving Ixodes ricinus and/or I. hexagonus ticks and the European hedgehog, Erinaceus europaeus, in suburban and urban areas in Switzerland. Folia Parasitologica 44, 309–314. Ginsberg, H.S. (ed.) (1993) Ecology and Environmental Management of Lyme Disease. Rutgers University Press, New Brunswick, New Jersey, 224 pp. Gray, J.S., Kahl, O., Janetzki, C. and Stein, J. (1992) Studies on the ecology of Lyme disease in a deer forest in Co. Galway, Ireland. Journal of Medical Entomology 29, 915–920. Kahl, O., Janetzki-Mittmann, C., Gray, J.S., Jonas, R., Stein, J. and de Boer, R.

44

O. Kahl et al.

(1998a) Risk of infection with Borrelia burgdorferi sensu lato for a host in relation to the duration of nymphal Ixodes ricinus feeding and the method of tick removal. Zentralblatt für Bakteriologie 287, 41–52. Kahl, O., Gern, L., Gray, J.S., Guy, E.C., Jongejan, F., Kirstein, F., Kurtenbach, K., Rijpkema, S.G.T. and Stanek, G. (1998b) Detection of Borrelia burgdorferi sensu lato in ticks: immunofluorescence assay versus polymerase chain reaction. Zentralblatt für Bakteriologie 287, 205–210. Kirstein, F. and Gray, J.S. (1996) A molecular marker for the identification of the zoonotic reservoirs of Lyme borreliosis by analysis of the blood meal in its European vector Ixodes ricinus. Applied and Environmental Microbiology 62, 4060–4065. Kurtenbach, K., Dizij, A., Seitz, H.M., Margos, G., Moter, S.E., Kramer, M.D., Wallich, R., Schaible, U.E. and Simon, M.M. (1994) Differential immune responses to Borrelia burgdorferi in European wild rodent species influence spirochete transmission to Ixodes ricinus L. (Acari, Ixodidae). Infection and Immunity 62, 5344–5352. Kurtenbach, K., Kampen, H., Dizij, A., Arndt, S., Seitz, H.M., Schaible, U.E. and Simon, M.M. (1995) Infestations of rodents with larval Ixodes ricinus (Acari, Ixodidae) is an important factor in the transmission cycle of Borrelia burgdorferi s.l. in German woodlands. Journal of Medical Entomology 32, 807–817. Labuda, M., Danielova, V., Jones, L.D. and Nuttall, P.A. (1993a) Amplification of tick-borne encephalitis virus infection during co-feeding of ticks. Medical and Veterinary Entomology 7, 339–342. Labuda, M., Jones, L.D., Williams, T., Danielova, V. and Nuttall, P.A. (1993b) Efficient transmission of tick-borne encephalitis virus between cofeeding ticks. Journal of Medical Entomology 30, 295–299. Lane, R.S. (1994) Competence of ticks as

vectors of microbial agents with an emphasis on Borrelia burgdorferi. In: Sonenshine, D.E. and Mather, T.N. (eds) Ecological Dynamics of Tick-borne Zoonoses. Oxford University Press, New York, pp. 45–67. Lane, R.S. and Keirans, J.E. (1999) Ixodes spinipalpis: a probable enzootic vector of Borrelia burgdorferi in California. In: Needham, G.R., Mitchell, R., Horn, D.J. and Welbourn, W.C. (eds) Acarology IX: Vol. 2, Symposia. Ohio Biological Survey, Columbus, Ohio, pp. 395–399. Lane, R.S. and Quistad, G.B. (1998) Borreliacidal factor in the blood of the western fence lizard (Sceloporus occidentalis). Journal of Parasitology 84, 29–34. Manelli, A., Kitron, U., Jones, C.J. and Slajchert, T.L. (1993) Ixodes dammini (Acari: Ixodidae) infestation on medium-sized mammals and blue jays in northwestern Illinois. Journal of Medical Entomology 30, 950–952. Mather, T.N., Wilson, M.L., Moore, S.I., Ribeiro, J.M.C. and Spielman, A. (1989) Comparing the relative potential of rodents as reservoirs of the Lyme disease spirochete (Borrelia burgdorferi). American Journal of Epidemiology 130, 143–150. Norris, D.E., Johnson, B.J.B., Piesman, J., Maupin, G.O., Clark, J.L. and Black, W.W. IV (1997) Culturing selects for specific genotypes of Borrelia burgdorferi in an enzootic cycle in Colorado. Journal of Clinical Microbiology 35, 2359–2364. Nuttall, P.A. and Labuda, M. (1994) Tickborne encephalitis subgroup. In: Sonenshine, D.E. and Mather, T.N. (eds) Ecological Dynamics of Tick-borne Zoonoses. Oxford University Press, New York, pp. 351–391. Pavlovsky, E.N. (1966) Natural Nidality of Transmissible Diseases with Special Reference to the Landscape Epidemiology of Zooanthroponoses. Levine, N.D. (ed.), University of Illinois Press, Urbana, 261 pp. (translated from Russian). Piesman, J. (1993) Standard system for

Ecological Research on B. burgdorferi

infecting ticks (Acari, Ixodidae) with the Lyme disease spirochete, Borrelia burgdorferi. Journal of Medical Entomology 30, 199–203. Piesman, J., Oliver, J.R. and Sinsky, R.J. (1990) Growth kinetics of the Lyme disease spirochete (Borrelia burgdorferi) in vector ticks (Ixodes dammini). American Journal of Tropical Medicine and Hygiene 42, 352–357. Randolph, S.E. and Craine, N.G. (1995) General framework for comparative quantitative studies on transmission of tick-borne diseases using Lyme borreliosis in Europe as an example. Journal of Medical Entomology 32, 765–777. Randolph, S.E. and Nuttall, P.A. (1994) Nearly right or precisely wrong? Natural versus laboratory studies of vector-borne diseases. Parasitology Today 10, 458–462. Randolph, S.E., Gern, L. and Nuttall, P.A. (1996) Co-feeding ticks: epidemiological significance for tick-borne pathogen transmission. Parasitology Today 12, 472–479. Randolph, S.E., Miklisová, D., Lysy, J., Rogers, D.J. and Labuda, M. (1999) Incidence from coincidence: patterns of tick infestations on rodents facilitate transmission of tick-borne encephalitis virus. Parasitology 118, 177–186. Randolph, S.E., Green, R.M., Peacey, M.F. and Rogers, D.J. (2000) Seasonal synchrony: the key to tick-borne encephalitis foci identified by satellite data. Parasitology 121, 15–23. Rijpkema, S.G.T., Molkenboer, M.J.C.H., Schouls, L.M., Jongejan, F. and Schellekens, J.F.P. (1995) Simultaneous detection and genotyping of three genomic groups of Borrelia burgdorferi s.l. in Dutch Ixodes ricinus ticks by characterization of the amplified intergenic spacer region between 5S and 23S rRNA genes. Journal of Clinical Microbiology 33, 3091–3095. Schwan, T.G., Burgdorfer, W. and Garon, C.F. (1988) Changes in infectivity and

45

plasmid profile of the Lyme disease spirochete, Borrelia burgdorferi, as a result of in vitro cultivation. Infection and Immunity 56, 1831–1836. Schwan, T.G., Piesman, J., Golde, W.T., Dolan, M.C. and Rosa, P.A. (1995) Induction of an outer surface protein on Borrelia burgdorferi during tick feeding. Proceedings of the National Academy of Sciences, USA 92, 2909–2913. de Silva, A., Telford, S.R., Brunet, L.R., Barthold, S.W. and Fikrig, E. (1996) Borrelia burgdorferi OspA is an arthropodspecific transmission-blocking Lyme disease vaccine. Journal of Experimental Medicine 183, 271–275. Sinsky, R.J. and Piesman, J. (1989) Ear punch biopsy method for detection and isolation of Borrelia burgdorferi from rodents. Journal of Clinical Microbiology 27, 1723–1727 (Erratum: (1989) Journal of Clinical Microbiology 27, 2401). Sonenshine, D.E. and Mather, T.N. (eds) (1994) Ecological Dynamics of Tick-borne Zoonoses. Oxford University Press, New York, 447 pp. Spielman, A., Levine, J.F. and Wilson, M.L. (1984) Vectorial capacity of North American Ixodes ticks. Yale Journal of Biology and Medicine 57, 507–513. Spielman, A., Telford, S. III and Pollack, R.J. (1993) The origins and course of the present outbreak of Lyme disease. In: Ginsberg, H.S. (ed.) Ecology and Environmental Management of Lyme Disease. Rutgers University Press, New Brunswick, New Jersey, pp. 83–96. Tälleklint, L. and Jaenson, T.G. (1994) Transmission of Borrelia burgdorferi s.l. from mammal reservoirs to the primary vector of Lyme Borreliosis, Ixodes ricinus (Acari, Ixodidae), in Sweden. Journal of Medical Entomology 31, 880–886. Telford, S.R. III, Mather, T.N., Moore, S.I., Wilson, M.L. and Spielman, A. (1988) Incompetence of deer as reservoirs of the Lyme disease spirochete. American Journal of Tropical Medicine and Hygiene 39, 105–109.

46

O. Kahl et al.

Wilson, M.L. (1994) Population ecology of tick vectors: interaction, measurement, and analysis. In: Sonenshine, D.E. and Mather, T.N. (eds) Ecological Dynamics of Tick-borne Zoonoses. Oxford University Press, New York, pp. 20–44. Wilson, M.L. (1998). Distribution and abundance of Ixodes scapularis (Acari, Ixodidae) in North America: ecological

processes and spatial analysis. Journal of Medical Entomology 35, 446–457. Wright, S.A., Lane, R.S. and Clover, J.R. (1998) Infestation of the southern alligator lizard (Squamata: Anguidae) by Ixodes pacificus (Acari: Ixodidae) and its susceptibility to Borrelia burgdorferi. Journal of Medical Entomology 35, 1044–1049.

Molecular and Cellular Biology of Borrelia burgdorferi sensu lato

3

Sven Bergström, Laila Noppa, Åsa Gylfe, and Yngve Östberg Department of Molecular Biology, Umeå University, SE-901 87 Umeå, Sweden

Introduction The genus Borrelia is a member of the spirochaete phylum, which is an ancient evolutionary branch only remotely related to Gram-negative and Gram-positive bacteria (Woese, 1987). The genus is divided into three major pathogenic groups of organisms, all transmitted by arthropods: the relapsing fever borreliae, Lyme disease (borreliosis) borreliae and the aetiological agent of avian spirochaetosis, Borrelia anserina. The Lyme disease borreliae of the Borrelia burgdorferi sensu lato (s.l.) species complex have been divided into at least 11 different genomic species (Table 3.1). Amongst these, the human pathogens include B. burgdorferi sensu stricto (s.s.), Borrelia afzelii, Borrelia garinii (Baranton et al., 1992; Marconi and Garon, 1992a; Canica et al., 1993) and possibly Borrelia bissettii (Picken et al., 1996; Strle et al., 1997). The strains designated to these genomic species are characterized by similar protein patterns, reactivity with monoclonal antibodies, amino acid sequence of outer surface proteins, 16S rRNA analysis, Southern blot and PCR analysis (Adam et al., 1991; Jonsson et al., 1992; Baranton et al., 1992), and 16S rRNA, flaB and 5S–23S rRNA intergenic spacer sequences (Wang, G. et al., 1999a). The distribution of Lyme borreliosis (LB) globally is closely tied to the presence of ticks of the Ixodes ricinus species complex in the northern hemisphere (see Eisen and Lane, Chapter 4). There are no confirmed cases of LB in the southern hemisphere apart from cases where it may have been acquired in Europe (Hudson et al., 1998), although B. garinii DNA has been amplified from the seabird tick Ixodes uriae in the southern hemisphere (Olsen et al., 1995). Members of the genus Borrelia have a very complex lifestyle as they are hostassociated spirochaetes that shuttle between haematophagous arthropods and © 2002 CAB International. Lyme Borreliosis: Biology, Epidemiology and Control (eds J. Gray, O. Kahl, R.S. Lane and G. Stanek)

47

48

S. Bergström et al.

Table 3.1. Borrelia burgdorferi sensu lato genospecies and their distribution. Borrelia species

Reference

Geographical distribution

B. burgdorferi s.s.

Baranton et al. (1992) Baranton et al. (1992)

Europe and North America Europe and parts of Asia

B. afzelii

Canica et al. (1993)

Europe and parts of Asia

B. valaisiana

Wang et al. (1997)

B. lusitaniae

Levy et al. (1993)

B. bissettii

Postic et al. (1998)

Central Europe, Ireland, Great Britain and the Netherlands Portugal, Tunisia, rarely in Central and Eastern Europe Slovenia and North America

B. japonica

Kawabata et al. (1993)

Japan

B. tanuki

Fukunaga et al. (1996a)

Japan

B. sinica

Masuzawa et al. (2001) Fukunaga et al. (1996a)

China

B. garinii

B. turdii

B. andersoni

Marconi et al. (1995)

Japan

North America

Reference (Baranton et al., 1998) (Hubálek and Halouzka, 1997; Baranton et al., 1998; Li et al., 1998) (Hubálek and Halouzka, 1997; Baranton et al., 1998; Li et al., 1998) (Hubálek and Halouzka, 1997; Kurtenbach et al., 1998) (Le Fleche et al., 1997; de Michelis et al., 2000) (Picken et al., 1996; Strle et al., 1997; Postic et al., 1998) (Kawabata et al., 1993; Fukunaga et al., 1996a; Masuzawa et al., 1996) (Kawabata et al., 1993; Fukunaga et al., 1996a; Masuzawa et al., 1996) (Masuzawa et al., 2001) (Kawabata et al., 1993; Fukunaga et al., 1996a; Masuzawa et al., 1996) (Marconi et al., 1995)

vertebrates. When the spirochaetes move from an arthropod to a vertebrate, they are changing from an environment with no antibody-based immune system to an environment where the immunological pressure is high; thus the biology of these organisms requires a high degree of adaptive capacity. In order to increase the probability of borreliae being transmitted by an infrequently feeding arthropod from one host to another, borreliae must persist in the host as long as possible. The antigens, which evoke immune reactions in the host, are mainly the surface molecules of the microorganism. Thus, a good way to escape the immune reaction is to change the molecules on the surface. This strategy is used by many microorganisms and may be accomplished in a number of ways, most of which involve DNA rearrangements (Borst, 1991). In addition, molecules on the bacterial surface mediate interaction between bacteria and host cells.

Molecular and Cellular Biology of B. burgdorferi

49

These are usually protein or glycolipid molecules. Adherence of B. burgdorferi s.l. to cultured epithelium, endothelium, differentiated neural cells, rat brain and glial cells has been shown (Garcia-Monco et al., 1989; Thomas and Comstock, 1989; Shoberg and Thomas, 1993; Peters and Benach, 1997). The ability of B. burgdorferi to infect different organ and tissue types might also be receptormediated, analogous to the tissue tropism described for Escherichia coli. Thus, it is obvious that the characteristics of the surface-exposed proteins and their ability to vary their expression or undergo specific changes is important for the biology of this organism.

Morphology, Shape and Membrane Structure Borreliae are thin, elongated, motile, wave-like bacteria (Fig. 3.1). A general feature of the spirochaetes, including B. burgdorferi s.l., is that they possess a fragile outer membrane surrounding the protoplasmic cylinder. This cylinder consists of a peptidoglycan layer, a cytoplasmic membrane (inner membrane) and the enclosed cytoplasmic contents (Johnson et al., 1984; Barbour and Hayes, 1986). The outer cell membrane is fluid and consists of 45–62% protein, 23–50% lipid and 3–4% carbohydrate (Barbour and Hayes, 1986). The borreliae have 7–11 bipolar flagella located in the periplasmic space where they are attached to the poles and wrapped around the cell cylinder, giving the bacterium its characteristic flat wave shape (Barbour and Hayes, 1986; Goldstein et al., 1994; Motaleb et al., 2000). Unlike other bacterial flagella, those of the spirochaetes are endoflagella, located in the periplasmic space between the protoplasmic cylinder and the outer cell membrane and attached to the poles of the bacterial cell. More than 100 polypeptides have been identified in the outer cell membrane of B. burgdorferi s.l. (Luft et al., 1989). The proteins most extensively studied are the outer surface proteins (Osps). To date, six Osps have been characterized, OspA–F. They are all surface-exposed and anchored by a lipid moiety to the fluid outer surface membrane (Brandt et al., 1990). A more complete understanding of the membrane constituents, their function and the overall membrane structure of B. burgdorferi s.l. will provide insight into the strategies that enable the bacterium to establish and maintain the infection during LB.

Flagella, motility and chemotaxis The periplasmic flagella (PFs), located in the periplasmic space between the cytoplasm–peptidoglycan layer and the outer membrane, have an essential role in cell morphology and motility (Sadziene et al., 1991; Goldstein et al., 1994, 1996; Motaleb et al., 2000). The distinct cell morphology of this bacterium is caused by the flagella that are wrapped around the protoplasmic cylinder (Burgdorfer et al., 1982; Hovind Hougen, 1984; Barbour and Hayes, 1986;

50

S. Bergström et al.

Fig. 3.1. Borrelia burgdorferi sensu lato stained by indirect immunofluorescence using the flagellin (FlaB) antibody H9724. (Photo courtesy of Björn Olsen.)

Goldstein et al., 1996). Each PF is attached at only one end and in Borrelia spp. they are long enough to overlap in the centre of the cell (Johnson et al., 1984). The genes encoding the flagellar proteins are flaA (Ge and Charon, 1997a), flaB (Wallich et al., 1990), flgE (Jwang et al., 1995), fliH and fliI (Ge et al., 1996), and the flgK operons (Ge et al., 1997), and are located on the chromosome. These genes have homologies to various genes in the flagellar apparatus of both animal and plant pathogens. This suggests that the flagellar apparatus and its protein export pathways are well conserved. Swimming B. burgdorferi s.l. cells have a flattened wave-like shape similar to that of eukaryotic flagella (Goldstein et al., 1994) A striking feature of B. burgdorferi s.l. is its capacity to swim efficiently in a

Molecular and Cellular Biology of B. burgdorferi

51

viscous medium such as connective tissue where other bacteria are slowed down or immobilized (Kimsey and Spielman, 1990; Goldstein et al., 1994). Motility in B. burgdorferi s.l. requires an environment similar to interstitial fluid (e.g. pH 7.6, 0.15 M NaCl and viscous) (Shi et al., 1998). The relative immobility of B. burgdorferi s.l. spirochaetes in less viscous media suggests active dissemination via the skin, as optimal mobility of borreliae is seen in collagen-like viscosity (Kimsey and Spielman, 1990). Furthermore, virulence studies on a spontaneously occurring non-flagellated, non-motile, mutant strain indicate that motility is important for the pathogenesis of this organism (Sadziene et al., 1991, 1996). The spirochaete motility apparatus is similar to that of other bacteria in that it has a filament, hook and basal body (Holt, 1978; Hovind Hougen, 1984; Barbour et al., 1986; Brahamsha and Greenberg, 1988; Charon et al., 1992). However, the PFs of spirochaetes differ from the flagella of most other bacteria as they are composed of two classes of proteins instead of one class (Cockayne et al., 1987; Norris et al., 1988; Koopman et al., 1992; Trueba et al., 1992; Rosey et al., 1995; Ge et al., 1998). These two classes are the flagellar outer sheath protein FlaA and the core protein FlaB. FlaA is unique to spirochaetes whereas FlaB shows homology to flagellar proteins of other bacteria (Magnarelli et al., 1987; Wallich et al., 1990). Until recently it was a general belief that the PFs of B. burgdorferi differed from those of other spirochaetes since the flagellar core protein FlaB was not thought to be surrounded by an outer sheath protein. However, using a modified extraction technique, Ge and Charon identified a FlaA homologue, showing 54–58% similarity to FlaA proteins from other spirochaetes (Ge and Charon, 1997a; Ge et al., 1998). In other spirochaetes, the FlaA and FlaB proteins are among the most abundant cell proteins (Koopman et al., 1992; Li et al., 1993; Norris, 1993). Depending on species, the PFs consist of one to two different FlaA proteins and two to four different FlaB proteins (Brahamsha and Greenberg, 1988; Norris et al., 1988; Charon et al., 1992; Koopman et al., 1992; Trueba et al., 1992; Li et al., 1993; Ge et al., 1998). In B. burgdorferi, the PFs are composed of only FlaA (39 kDa) and FlaB (41 kDa), where FlaB is one of the most abundant proteins and FlaA is expressed in lower amounts. Chemotaxis studies have revealed that rabbit serum acts as an attractant while ethanol and butanol act as repellents (Shi et al., 1998). Interestingly, B. burgdorferi does not exhibit chemotaxis towards common sugars or amino acids (Shi et al., 1998). It has also been noted that spirochaetes grown in tissue chambers implanted in rats (host-adapted bacteria) show more vigorous motility compared with spirochaetes grown in vitro (Akins et al., 1998). The chemotaxis genes cheA, cheE (Trueba et al., 1997), cheW and cheY (Fraser et al., 1997; Ge and Charon, 1997b; Trueba et al., 1997) have been identified in B. burgdorferi s.l. In E. coli, chemotaxis is achieved by regulating the direction of rotation of the flagellar bundle. In B. burgdorferi and other spirochaetes, this is not easily achieved since they have two flagellar bundles, one at each cell end. In order to switch the direction of swimming, they have to rotate in different directions and then switch synchronously. This type of motility is observed as smooth swimming, used when an attractant is present. Another type of observed swimming pattern, a flexing

52

S. Bergström et al.

movement, is observed in the presence of repellents. The flexing movement is the result of two bundles rotating in the same direction (Charon et al., 1992; Goldstein et al., 1994; Shi et al., 1998).

Structure of the outer membrane Since B. burgdorferi s.l. possesses both an outer membrane and a cytoplasmic membrane, the surface of this bacterium is analogous to enteric Gram-negative bacteria. However, its membrane composition differs from that of Gram-negative bacteria and has several distinct features (Fig. 3.2). For example, B. burgdorferi s.l. has an extraordinarily high abundance of membrane proteins covalently modified with lipids (Brandt et al., 1990; Fraser et al., 1997). Another characteristic is the absence of phosphatidylethanolamine (Belisle et al., 1994) and lipopolysaccharide (LPS) (Takayama et al., 1987), and the presence of glycolipid antigens other than LPS (Eiffert et al., 1991; Wheeler et al., 1993; Belisle et al., 1994). Glycolipids represent about 50% of the total lipids and comprise only galactose as monosaccharide constituents (Berg, 2001; Hossain et al., 2001). The existence of LPS on the surface of B. burgdorferi s.l. has been controversial, although most evidence currently points to its absence (Beck et al., 1985; Takayama et al., 1987; Cinco et al., 1991; Schoenfeld et al., 1992; de Souza et

Lipoproteins LPS

OM

P IM B. burgdorferi

E. coli

Fig. 3.2. Comparison of Borrelia burgdorferi sensu lato and Escherichia coli cell envelopes. IM, inner membrane; LPS, lipopolysaccharide; OM, outer membrane; P, protein.

Molecular and Cellular Biology of B. burgdorferi

53

al., 1992). The B. burgdorferi outer membrane contains a relatively low density of transmembrane-spanning proteins as determined by freeze–fracture electron microscopy studies (Walker et al., 1991; Radolf et al., 1994; Jones et al., 1995). This may explain why B. burgdorferi is more susceptible to disruption by routine physical manipulations such as centrifugation and resuspension and is more susceptible to detergents compared with Gram-negative bacteria. It is very likely that the unusual surface of this bacterium contributes to the unique properties that B. burgdorferi s.l. possesses, e.g. the ability to survive in both the mammalian host and the tick and to evade the mammalian immune system. Thus, a comprehensive analysis of the membrane architecture and constituents would yield insight into the unique physiological mechanisms by which the bacterium survives in diverse environments and may also elucidate pathogenic mechanisms operative during LB.

Blebs B. burgdorferi s.l. is capable of producing extracellular vesicles (blebs) when grown in laboratory culture media (Barbour and Hayes, 1986). From earlier studies on the aetiologic agent of relapsing fever, it is evident that a number of different chemical and physical agents can disturb the bacteria, i.e. when the borrelial cells are treated with penicillin, freeze–thawing, prolonged cultivation, or the addition of a specific antibody together with a complement source, the spirochaetes form outer membrane blebs which detach from the parent organism (Barbour and Hayes, 1986). In the case of B. burgdorferi, the exposure to stress from the mammalian milieu, even in the absence of anti-B. burgdorferi antibodies, is enough to trigger the production of membrane-bound and possibly secreted blebs (Shoberg and Thomas, 1995). All borrelial blebs appear to contain an outer membrane, while some also contain cytoplasmic membranes (Radolf et al., 1994). However, all vesicle preparations made to date have no detectable quantities of the periplasmic 41 kDa flagellin, investigated by Western blotting (Dorward et al., 1991; Shoberg and Thomas, 1995). These blebs contain plasmid DNA (Garon et al., 1989), including osp DNA (Persing et al., 1994), and in their membranes the OspA and OspB proteins and another low molecular weight lipoprotein, Lp6.6, have been found (Dorward et al., 1991; Katona et al., 1992; Lahdenne et al., 1997). The bleb membranes are also thought to contain the borrelial adhesin for human endothelial cells since they are capable of binding to these cells (Shoberg and Thomas, 1993). As an anecdote, in the first half of the 20th century an alternative viewpoint was that these large blebs or ‘gemmae’ represented another stage of the life cycle of spirochaetes and this encouraged the early classification of borreliae as protozoans (Barbour and Hayes, 1986). The recently described phenomenon of cyst formation is another sign of disturbance of cell architecture. Cyst formation involves changes in membrane integrity and cell shape (Alban et al., 2000). It is apparently a starvation response, which can be induced by cultivation in serum-free BSKII medium. It is a slow

54

S. Bergström et al.

but active process requiring protein synthesis and is reversible in up to 53% of the cells. It is not known whether cyst formation is important for persistence of borreliae in the host and the explanation for survival during antibiotic therapy.

The outer membrane proteins There is much evidence that B. burgdorferi s.l. contains an unusually high number of proteins covalently modified by lipids (e.g. lipoproteins). The recently published genome sequence of B. burgdorferi s.s. revealed 132 open reading frames (corresponding to about 4.9% of the coding sequence) that contained the signal sequence characteristic for lipoproteins and 32 additional open reading frames that had other similarities to lipoproteins (Hayashi and Wu, 1990; Fraser et al., 1997; Casjens et al., 2000). Typically, proteins involved in interaction with the host are encoded on plasmids and, on the B. burgdorferi s.l. plasmids, putative lipoproteins make up at least 14.5% of the open reading frames (Casjens et al., 2000). The complete sequence of other related bacterial genomes revealed lower proportions of putative lipoproteins. For example, in Treponema pallidum and Helicobacter pylori, only 2.1% and 1.3%, respectively, of the coding sequences were potential lipoproteins (Tomb et al., 1997; Fraser et al., 1998; Casjens et al., 2000). Direct evidence for a large number of lipoproteins in B. burgdorferi was obtained by metabolic labelling of in vitro-growing cultures of B. burgdorferi with [3H]-palmitic acid, which revealed an abundance of metabolically labelled lipoproteins (Brandt et al., 1990). The function of the lipid moiety attached to lipoproteins is to anchor these polypeptides to the appropriate membrane. These modifications can then ensure membrane localization of otherwise hydrophilic proteins. Illustrating this, three T. pallidum lipoproteins are hydrophilic when synthesized in an in vitro transcription–translation system where acylation does not occur (Chamberlain et al., 1989; Swancutt et al., 1989) and a recombinant nonlipidated form of OspA is soluble in aqueous solutions (Dunn et al., 1990). The modification of lipoproteins occurs in several steps. First, the signal sequence LXYC specifies the addition of diacylglycerol in a thioether linkage to the sulphur of the cysteine side chain. Then signal peptidase II cleaves at the amino side of the cysteine residue and a third fatty acid is added in an amide linkage to the new amino terminus (Hayashi and Wu, 1990). The enzymes responsible for these reactions were first discovered in E. coli (Sankaran et al., 1995). It is likely that B. burgdorferi lipoproteins are generated in a similar fashion since B. burgdorferi encodes enzymes highly homologous to the E. coli enzymes that are required for lipoprotein maturation (Fraser et al., 1997). It appears that the immunopotentiating activities of bacterial lipoproteins are mostly conferred to their lipid constituents since the lipid moiety can elicit an inflammatory response (Bessler et al., 1985; Lex et al., 1986; Hoffmann et al., 1988; Deres et al., 1989; Erdile et al., 1993). Further studies have revealed that synthetic lipohexapeptides of varying peptide composition, but sharing a tripalmitoyl-cysteine residue, were equally effective (Radolf et al., 1995b).

Molecular and Cellular Biology of B. burgdorferi

55

However, non-lipidated recombinant OspA is still recognized by antibodies from LB patients (Dunn et al., 1990). Prokaryotic lipoproteins are secreted through the Sec-dependent general secretory pathway and, in general, the periplasmic space is the final destination for their polypeptide portions (Pugsley, 1993). This hypothesis is supported by the fact that lipoproteins are rarely surface-exposed in Gram-negative bacteria and that lipidated OspA expressed in E. coli is associated with the cytoplasmic membrane (Dunn et al., 1990; Pugsley, 1993). In T. pallidum, the agent for venereal syphilis, all lipoproteins appear to be located beneath the surface (Cox et al., 1992; Radolf, 1994). In B. burgdorferi s.l. there is evidence for outer-membrane association of many lipoproteins, and in some cases lipoproteins appear to be located in both the outer and inner (cytoplasmic) membranes. Many borrelial proteins are surfaceexposed based on experimental evidence, e.g. proteinase K accessibility assays of intact cells and immunogold labelling (Barbour et al., 1984; Bunikis et al., 1995; Carroll et al., 2001; Noppa et al., 2001). Except for minor differences in relative abundance, the lipoprotein constituents in outer membranes from borreliae labelled with [3H]-palmitic acid were essentially identical to those of whole cells (Radolf et al., 1995a). However, not all lipoproteins in B. burgdorferi s.l. are located in the outer membrane since the protoplasmic cylinders contained lipoproteins not observed in the outer membranes (Radolf et al., 1995a). Several studies suggest that the predominant location of the lipoproteins OspA and OspB is in the periplasmic space, but these proteins are also found in the fluid outer membrane (Brusca et al., 1991; Bledsoe et al., 1994; Radolf et al., 1995a; Skare et al., 1995; Carroll and Gherardini, 1996; Cox et al., 1996). In contrast to the highly abundant lipoproteins in the borrelial outer membrane, the abundance of transmembrane-spanning proteins in the outer-membrane proteins of B. burgdorferi s.l. is very low compared with that of Gram-negative bacteria. However, when compared with another spirochaete, T. pallidum, the membrane-spanning proteins are approximately tenfold more abundant in B. burgdorferi s.l., as determined by freeze–fracture electron microscopy studies (Walker et al., 1991; Radolf et al., 1994).

The major Osp lipoproteins Several lipoproteins have been characterized in B. burgdorferi s.l. Many of these proteins are called outer surface proteins, Osps, and most of these proteins do not have any significant sequence or structural homology to proteins of known function. When various B. burgdorferi s.l. isolates are compared, the Osps show considerable heterogeneity within and among the different species. Both OspA and OspB are encoded by the same operon on a 49 kb linear plasmid (Barbour and Garon, 1988; Bergström et al., 1989). The first gene that was cloned and sequenced in B. burgdorferi, ospA, encodes the most studied protein in these bacteria (Howe et al., 1986; Bergström et al., 1989). In an attempt

56

S. Bergström et al.

to reveal the function of OspA, the crystal structure of a complex of recombinant unlipidated OspA and the Fab fragment of an anti-OspA monoclonal antibody was elucidated (Li et al., 1997). The structure was not consistent with that of an enzyme but it had a hydrophobic cavity buried in a positively charged cleft in the carboxyterminal domain, which might be a binding site for an unknown ligand. This presence of conserved charged groups in a pronounced cleft is typical of proteins that bind small or linear polymeric ligands. OspA could thus recognize either a soluble or host surface molecule and possibly transmit this information to another membrane component, suggesting that OspA could act as a receptor (Li et al., 1997). The structure of OspA also indicated that the conserved NH2-terminus could interact with another membrane protein (Li et al., 1997). OspA has been shown to be expressed by the spirochaete in the unfed tick’s midgut and mediate adhesion to the midgut epithelium (Pal et al., 2000). This preferential expression in the midgut suggests that OspA may have an important function in the vector. Thus, this OspA-mediated attachment to the tick midgut epithelium provides a possible mechanism for how stage-specific protein expression can contribute to pathogenesis during the natural cycle of borrelial spirochaetes (Pal et al., 2000). OspA and OspB have, in a large number of in vivo and in vitro experiments, been shown to play important roles in the pathogenesis of LB. Studies have shown that these two major proteins are involved in adherence to and penetration of mammalian host cells. Thus, anti-OspA monoclonal antibodies inhibit adhesion of B. burgdorferi to endothelial cell surfaces, and during subculturing in BSKII medium the spirochaetes lose OspB and their infectivity (Schwan and Burgdorfer, 1987; Schwan et al., 1988; Comstock et al., 1993). Furthermore, mutants lacking OspB do not penetrate into endothelial cells and their infectivity is from 30- to 300-fold lower than that of wild-type spirochaetes (Sadziene et al., 1993b). Development of what has been termed treatment-resistant Lyme arthritis has been shown to be associated with particular major histocompatibility complex class II alleles and immunoreactivity to the OspA protein (Chen et al., 1999). Most of the patients with treatment-resistant Lyme arthritis have HLADRB1*0401 or related alleles, and the severity and duration of their arthritis correlate with cellular and humoral immune responses to the B. burgdorferi s.s. OspA protein. In a gene bank homology search, one human protein, the lymphocyte function-associated antigen-1 (hLFA-1), which had sequence homology with OspA165–173 and predicted binding in the DRB1*0401 molecule, was identified (Gross et al., 1998). Only individuals with treatment-resistant Lyme arthritis, in contrast to other forms of arthritis, generated these responses to OspA, hLFA-1 and the highly related peptide epitopes. From these results, Steere and co-workers proposed a model of molecular mimicry affecting genetically susceptible individuals to explain the autoimmune mechanisms in antibiotic treatment-resistant Lyme arthritis (Steere et al., 2001; Trollmo et al., 2001). OspC is homologous to proteins in both the relapsing fever species Borrelia hermsii and other Borrelia spp. Because of this, these proteins were suggested as

Molecular and Cellular Biology of B. burgdorferi

57

defining a genus-wide protein family (Carter et al., 1994). OspC is very heterogeneous and a large degree of molecular polymorphism of the ospC gene has been shown when analysing the different Borrelia genospecies (Jauris-Heipke et al., 1995; Livey et al., 1995). So far 19 major OspC groups have been identified (Wang, I.N. et al., 1999). Molecular analysis of the ospC sequences from different Borrelia genospecies suggests that recombination between different ospC alleles can occur frequently and that this genetic exchange is mediated by lateral transfer between ospC sequences. This can be achieved either by lateral transfer of ospC genes from the same genospecies or between different Borrelia spp. (Livey et al., 1995; Wang, G. et al., 1999b). There is an inverse relationship between the expression of OspA and OspC in vitro: strains with low expression of OspC have strong expression of OspA and vice versa (Margolis and Rosa, 1993). During tick feeding, the spirochaete alters its expression of Osps. In non-engorged ticks, the spirochaete expresses and produces OspA and probably OspB on the surface. Furthermore, OspA has been shown to mediate adhesion to the midgut epithelium in the unfed tick’s midgut (Pal et al., 2000). The production of OspA and OspB proteins decreases at the time of feeding, and instead the spirochaete promotes production of OspC (Schwan et al., 1995; Ohnishi et al., 2001). OspC expression is induced 36–48 h into the blood meal and probably mediates the escape of the spirochaete from the tick midgut via the haemolymph to the salivary glands from where it enters the new vertebrate host (Anderson et al., 1989; Schwan et al., 1995; Schwan and Piesman, 2000; Seshu and Skare, 2000; Ohnishi et al., 2001). The OspC protein is encoded on a 27 kb circular plasmid. Recent studies have suggested that the expression of the ospC gene can be influenced, in an inverse relationship, by the expression of the ospA and ospB genes (Sadziene et al., 1993a; Jonsson and Bergström, 1995; Schwan et al., 1995; Schwan and Piesman, 2000). Antibodies against OspA in the blood of vertebrate hosts kill spirochaetes in the tick midgut before dispersal to the salivary glands and thereby block transmission to the vertebrate host (de Silva et al., 1996, 1999). Antibodies against OspC are also efficient, probably blocking the transmission by preventing migration of spirochaetes to the salivary glands (Gilmore and Piesman, 2000). Recombinant OspA is an efficient vaccine against LB, and immunization with OspC confers protective immunity in animal studies (Gilmore et al., 1996; Sigal et al., 1998; Steere et al., 1998). However, the use of these vaccines may be limited by the apparent variability of these proteins (Wilske et al., 1996). OspD is an abundant 28 kDa surface-exposed component of B. burgdorferi s.l. (Norris et al., 1992). The ospD gene is located on a 38 kb linear plasmid present in most low-passage strains of B. burgdorferi. Molecular cloning and sequence analysis of the ospD gene locus revealed an open reading frame encoding a 28,436 Da polypeptide with a putative signal peptidase II leader sequence (Norris et al., 1992). Sequence analysis has demonstrated the presence of various numbers of 17 bp repeated sequences in the upstream control (promoter) region of the gene (Norris et al., 1992; Marconi et al., 1994). However, only one transcription initiation site was active as determined by primer extension analysis (Norris et al., 1992).

58

S. Bergström et al.

OspE and OspF are the founding members of a protein family known as the OspEF-related proteins (Erps) (Lam et al., 1994b; Stevenson et al., 1996). Individual bacteria carry multiple erp operons, each located on a different plasmid of the cp32 family (Stevenson et al., 1996; Casjens et al., 1997a; Akins et al., 1999; Stevenson et al., 2000a). Clonal B. burgdorferi with as many as nine different erp loci per cell have been characterized (Stevenson et al., 1996, 1998; Casjens et al., 1997a; Akins et al., 1999; El-Hage et al., 1999; Casjens et al., 2000). Since several Erps are immunogenic, these lipoproteins are surface-exposed and may therefore be important for the virulence of B. burgdorferi. Surface localization of these infection-associated proteins indicates the potential for interactions of Erps with vertebrate tissues. At least some Erps can bind complement factor H, suggesting that these proteins might be important for the borreliae during mammalian infection by inhibition of complement activation. Recombination of the genes encoding the Erp antigens might contribute to the evasion of the mammalian immune response and could play roles in the establishment and persistence of B. burgdorferi s.l. infections in mammalian hosts (Hellwage et al., 2001). All LB spirochaetes contain erp genes, but these genes have not been found in other species of the genus Borrelia, indicating that Erps perform a function unique to the biology of LB borreliae (Stevenson et al., 2000a,b). Recently, an immunodominant Osp of B. afzelii, Osp17, was identified (JaurisHeipke et al., 1999). Moreover, the B. burgdorferi genome sequence defined several hypothetical lipoproteins, which might increase the number of possible Osp proteins in the future.

The decorin-binding proteins Microbial adhesion to and colonization of host tissues are an important event in the bacterial infection process. Most pathogenic bacteria express adhesins that can participate in attachment mechanisms that might result in tissue colonization. Spirochaetes injected in the skin by an ixodid tick are found associated with collagen fibres in the extracellular matrix (ECM). However, B. burgdorferi s.l. do not attach directly to collagen but to decorin, a small leucine-rich proteoglycan (SLRP) that is associated with collagen fibres (Guo et al., 1995, 1998). Decorin is a collagen-associated ECM proteoglycan found in the skin (the site of entry for the spirochaete) and in many other tissues. There are two borrelial-encoded adhesins that can bind to decorin, DbpA and DbpB (decorin-binding protein) (Guo et al., 1995; Hanson et al., 1998). DbpA and DbpB were also isolated as antigens that were recognized by sera from mice infected with B. burgdorferi s.s. (Feng et al., 1998; Hagman et al., 1998). Both DbpA and DbpB are outer membrane-associated and may function as adhesins to facilitate bacterial adherence to mammalian tissue. However, alternative adhesion mechanisms may also be used by the spirochaetes depending on the tissue target and stage of dissemination (Duray, 1992; van Mierlo et al., 1993). To further investigate the role of Dbps in the pathogenesis of LB, a decorin-deficient (Dcn–/–) mouse was devel-

Molecular and Cellular Biology of B. burgdorferi

59

oped and used in experimental LB (Brown et al., 2001). In that study it was demonstrated that Dcn–/– mice were more resistant to LB, which suggests that decorin could be a limiting factor as a substrate for adherence of B. burgdorferi s.l. Interestingly, this resistance to B. burgdorferi s.l. in Dcn–/– could be reduced by increasing the infection dose, which indicates that additional adhesion ligands can participate in the disease process in vivo (Brown et al., 2001). Another adhesin, the fibronectin-binding protein BBK32, has been identified in all LB genospecies (Probert and Johnson, 1998). Molecular analysis of the bbk32 gene and gene product defined its fibronectin-binding domain, which shared sequence homology with a fibronectin-binding peptide in Streptococcus pyogenes, indicating a common mechanism of adhesion for B. burgdorferi s.l. and S. pyogenes (Probert et al., 2001).

The oligopeptide permease (Opp) Homologues of all five subunits of the peptide-binding component OppA (oppAI–V) as well as the other components of oligopeptide permease (OppB, OppC, OppD, OppF) were identified and characterized in B. burgdorferi s.l. (Das et al., 1996; Bono et al., 1998; Kornacki and Oliver, 1998). It is not known if oligopeptide permease has a sensing capability in B. burgdorferi s.l. similar to other bacteria although an interesting feature to note is that Opp-like proteins are involved in expression of virulence determinants in other bacteria (Payne and Gilvarg, 1968; Goodell and Higgins, 1987; Finlay and Falkow, 1989; Gominet et al., 2001). The Opp system is particularly interesting in B. burgdorferi s.l. since it was evident from the genome sequence analysis that only a rudimentary machinery for synthesis or transport of amino acids exists. In fact, only a single intact peptide transporter operon could be found. From the genome analysis of B. burgdorferi s.l. it is also apparent that the opp operon differs from that of E. coli as it has three separate substrate-binding proteins, OppA-1, 2 and 3. Interestingly, all of the B. burgdorferi s.l. OppA proteins were found to complement E. coli OppBCDF to form a functional peptide transport system for nutritional purposes (Lin et al., 2001).

The vmp-like sequence, Vls, of B. burgdorferi s.l. To increase their chances of transmission by an arthropod from one host to another, borreliae persist in the host for as long as possible. Persistence can be achieved by a mechanism called antigenic variation, and this mechanism has been studied for several years in relapsing fever borreliosis where new serotypes appear during relapses as a change in novel surface proteins called variable major proteins (Vmps). In the relapsing fever spirochaete B. hermsii, at least 26 different surface-exposed lipoproteins encoded by genes situated on linear plasmids are alternately expressed and presented, one at a time, on the bacterial surface

60

S. Bergström et al.

(Barbour, 1990, 1993; Burman et al., 1990; Kitten and Barbour, 1990; Barbour et al., 1991a). Only one vmp gene, located at the expression site, is expressed whereas the other genes are kept silent at storage sites. Antigenic variation occurs when a silent vmp gene replaces the expressed vmp gene by homologous recombination (Plasterk et al., 1985; Barbour et al., 1991b; Restrepo and Barbour, 1994; Restrepo et al., 1994). In B. burgdorferi s.s a 28 kb linear plasmid was found to have a vmp-like sequence (vls locus) that resembles the Vmp system in B. hermsii (Zhang et al., 1997). Fifteen non-expressed vls cassettes can be recombined into a functional and expressed vlsE region resulting in antigenic variation of the expressed lipoprotein (Zhang et al., 1997) and involving partial gene conversion of the expressed vls gene by parts of the vls genes at silent loci in a unidirectional process (Zhang and Norris, 1998a,b). The mechanism of Vls antigenic variation in LB Borrelia spp. is similar to the mechanism that is responsible for the recombination between the expressed pilin gene pilE and silent copies of pilS in Neisseria gonorrhoeae (Hagblom et al., 1985; Haas and Meyer, 1986; Zhang et al., 1992). VlsE contains two invariable domains and one variable domain that includes six variable and six invariable regions. Among molecules that undergo antigenic variation, VlsE is unusual in that more than 75% of its primary structure is invariable (Liang and Philipp, 1999). The relevance of the antigenic variation of VlsE in the pathogenesis of B. burgdorferi s.l. remains unclear. Even though there is a correlation between the presence of the plasmid that encodes the Vls lipoprotein and infectivity, the Vls proteins appear to be a minor constituent of the bacterial surface. However, it is likely that the antigenic variation expressed by the vls gene locus of B. burgdorferi s.l. is used in order to persist in mammals and to evade the host immune response. Little is known about the signals that trigger recombination events at the vls locus. Recombination occurs in vivo as early as 4 days after experimental infection of mice but not in vitro, suggesting that the mammalian host provides the signal for vls recombination (Zhang and Norris, 1998a). The role of the mammalian host is important since the vlsE recombination rates by B. burgdorferi s.l. were found to be lower in IFN-γ R-deficient mice than in control animals. These results suggest that the murine immune response can promote the in vivo adaptation of B. burgdorferi s.l. (Anguita et al., 2001). In addition to the importance of VlsE in the immune evasion of B. burgdorferi s.l., the ospE family of genes is able to undergo mutational changes during infection, leading to the generation of antigenically distinct OspE protein variants. This adds OspE to the repertoire of antigens that can be important for adaptation and immune evasion during LB (Sung et al., 2000).

Integral membrane proteins Bacteria other than B. burgdorferi s.l. commonly ensure correct localization of the membrane proteins by including transmembrane-spanning domains in their proteins. Proteins containing such domains appear to be rare in B. burgdorferi s.l.

Molecular and Cellular Biology of B. burgdorferi

61

This is based on the observation that biosynthetic labelling of borrelia cultures with [14C]-amino acids identified only a few amphiphilic polypeptides that do not co-migrate with lipoproteins (Brandt et al., 1990). It was also shown that these putative transmembrane proteins were in relatively low abundance when visualized by freeze–fracture electron microscopy (Walker et al., 1991; Radolf et al., 1994). Virtually none of the proteins were recognized by sera from patients with chronic LB (Brandt et al., 1990). Thus, B. burgdorferi s.l. contains two classes of integral membrane proteins: abundant lipoproteins and rare transmembranespanning proteins. Three characterized membrane proteins have been predicted to contain transmembrane-spanning domains, P66 (Oms66), Oms28 (Oms refers to outer membrane spanning) and P13 (Bunikis et al., 1995; Skare et al., 1996, 1997; Noppa et al., 2001). P66 and Oms28 have also been shown to exhibit porin activities (Skare et al., 1996, 1997). In Gram-negative bacteria, porins are known to form large water-filled channels that allow the diffusion of hydrophilic molecules into the periplasmic space, thus facilitating transport of molecules across the bacterial membrane (Cowan et al., 1992). The third protein, P13, contains transmembrane-spanning domains and was shown to be surface-exposed by immunofluorescence assays, immunoelectron microscopy and protease sensitivity assays (Sadziene et al., 1995; Noppa et al., 2001). The deduced sequence of the P13 peptide revealed possible signal peptidase type I cleavage sites and the computer analysis predicted that P13 has three transmembrane-spanning domains. Mass spectrometry, in vitro translation and N- and C-terminal amino acid sequencing analyses have indicated that P13 is very unusual as it is posttranslationally processed at both ends and modified by an unknown mechanism (Noppa et al., 2001).

Other surface-exposed membrane proteins We do not intend to cover all aspects and importance of every membrane protein of B. burgdorferi s.l. However, we will add some comments below on the characteristics of the outer surface-associated proteins and protein families P22, P35, Bmp, Mlp, RevA and Epp. Many of these proteins are probably important both as virulence determinants and as possible antigens for prophylactic and diagnostic use. P22 P22 is a 21.8 kDa protein encoded from the linear chromosome and is homologous to the IpLA7 protein (98.5% homology) (Simpson et al., 1991a; Wallich et al., 1993). The P22 protein is a surface-exposed lipoprotein, as demonstrated by [3H]-palmitate and immunogold labelling (Lam et al., 1994a).

62

S. Bergström et al.

P35 This is yet another B. burgdorferi s.l. lipoprotein that is recognized by sera from B. burgdorferi s.l.-infected mammals. Molecular analysis revealed that the p35 gene was selectively expressed in vivo and is upregulated or initiated during the postlogarithmic bacterial growth phase in vitro (Indest et al., 1997; Indest and Philipp, 2000). Bmp (P39) BmpA, BmpB, BmpC and BmpD constitute a family of homologous lipoproteins that are immunogenic during infection in vertebrate hosts. The genetic organization of the bmp region has been characterized for a variety of B. burgdorferi s.l. strains by Southern hybridization, PCR amplification and DNA sequencing (Simpson et al., 1991b; Gorbacheva et al., 2000). Mlp Mlp is the designation of a B. burgdorferi s.l. family of multicopy lipoproteins encoded in at least seven versions on the circular cp32 and cp18 plasmids (Porcella et al., 1996; Yang et al., 1999). The mlp transcription has an ospC-like temperature regulation as the mlp genes are not expressed at low temperature but expression is increased when shifted to higher temperatures (Porcella et al., 2000; Yang et al., 2000). RevA RevA is another B. burgdorferi s.l. surface-exposed protein that is expressed at high temperature in the mammalian host. The level of RevA expression was also found to be dependent upon the pH of the culture (Carroll et al., 2001). EppA The exported plasmid protein A (eppA) genes are only expressed during infection in the mammalian host (Champion et al., 1994). The eppA genes are located on either the plasmid cp9-1 or cp9-2 (Miller et al., 2000). The widespread occurrence of eppA genes and 9 kb circular plasmids suggests that they contribute to the survival of B. burgdorferi s.l. in nature, and possibly to the pathogenesis of LB (Champion et al., 1994; Miller et al., 2000).

Genome and Genome Organization B. burgdorferi s.l. has a low guanine and cytosine content of 28–30.5 mole % (Schmid et al., 1984; Bergström et al., 1989; Fraser et al., 1997). The low DNA homology and the low guanine and cytosine content separate the genus Borrelia from the genera of Treponema and Leptospira (Hyde and Johnson, 1984).

Molecular and Cellular Biology of B. burgdorferi

63

The genetic organization of Borrelia spp. is unusual among prokaryotes in the sense that they contain a small linear chromosome approximately 1 Mb in size (Baril et al., 1989; Ferdows and Barbour, 1989; Bergström et al., 1992; Davidson et al., 1992; Casjens and Huang, 1993; Fraser et al., 1997) and numerous linear and circular plasmids (Barbour and Garon, 1987; Barbour, 1988; Casjens et al., 1997a). Most of the genome of the B. burgdorferi s.s. strain B31 was published in 1997 (Fraser et al., 1997) and it was completed in 2000 (Casjens et al., 2000). The sequencing of the B. burgdorferi s.s. B31 genome revealed a total genome size of 1.52 Mbp: 910,725 bp on the linear chromosome, and 611 kbp divided on 12 linear and nine circular plasmids (Fig. 3.3). There were only 1283 putative genes

Chromosome: 0.91 Mb cp9

cp26 cp32 × 7

lp5

Plasmids (total): 611 kb lp17 lp21 lp25 lp28 (1–4) lp36 lp38 lp54 lp56

Fig. 3.3. Schematic representation of the total genome of Borrelia burgdorferi sensu stricto B31.

64

S. Bergström et al.

(Fraser et al., 1997), compared with the 4405 genes present in E. coli. The small number of genes in B. burgdorferi s.s is similar to the number of genes in the related spirochaete T. pallidum (Fraser et al., 1998) and the obligate intracellular bacterium Mycoplasma genitalium (Fraser et al., 1995). As in T. pallidum and M. genitalium, a number of genes encoding cellular biosynthetic enzymes are missing in B. burgdorferi s.s. B31. Since Mycoplasma spp. and spirochaetes are only distantly related, the limited metabolic capacity probably reflects convergent evolution by the loss of genes, allowed by the parasitic lifestyle of these bacteria. A linear chromosome has been described in only two other eubacteria, Streptomyces lividans (Lin et al., 1993) and Agrobacterium tumefaciens (Allardet-Servent et al., 1993), while linear plasmids have been observed in Streptomyces spp., Rhodococcus fascians and Thiobacillus versutus (Kinashi et al., 1987; Wlodarzyk and Nowicka, 1988; Crespi et al., 1992). The vast majority of bacterial species have a circular chromosome and circular plasmids. The ends of the Borrelia spp. chromosome and linear plasmids are covalently closed loops similar to eukaryotic telomeres (Hinnebusch et al., 1990; Casjens et al., 1997b). Replicons with covalently closed hairpin ends have also been described in several other organisms including poxviruses, African swine fever virus, Chlorella viruses, the E. coli N15 prophage, certain mitochondrial DNAs, mitochondrial plasmids and plastid DNAs (Meinhardt et al., 1997; Nosek et al., 1998; Rybchin and Svarchevsky, 1999; Volff and Altenbuchner, 2000). The problem of the replication of B. burgdorferi s.l. DNA molecules has been discussed ever since the discovery of its linear DNA replicons. The mechanisms by which the hairpin telomeres are processed during replication have now been investigated by Chaconas et al. (2001). They showed that a synthetic 140 bp sequence, having the predicted structure of a replicated telomere, could function as a viable substrate for telomere resolution in vivo, and as such was sufficient to convert a circular replicon to a linear form. They suggested that the final step in the replication of linear B. burgdorferi s.l. replicons is a site-specific DNA breakage and reunion event to regenerate covalently closed hairpin ends (Chaconas et al., 2001).

The linear chromosome (the maxi chromosome) The genomic sequence of the B. burgdorferi linear chromosome contains 853 chromosomal genes encoding proteins involved in fundamental processes such as replication, transcription, translation, energy metabolism and transport across the membranes. The average size of the open reading frames is 992 bp, and 93% of the genome is within coding sequences. The GC content of only 28.6% of the B. burgdorferi s.s. chromosome has been confirmed and all 61 possible codon triplets are used, with a marked bias towards the AU-rich triplets (Hyde and Johnson, 1984; Johnson et al., 1984; Schmid et al., 1984; Fraser et al., 1997).

Molecular and Cellular Biology of B. burgdorferi

65

Linear plasmids (mini chromosomes) and circular plasmids The genome sequence of the entire B. burgdorferi s.s. B31 isolate MI was partially published in 1997 and completed in 2000. From those studies it was shown that B. burgdorferi s.s. B31 carries 21 extrachromosomal DNA elements, which is the largest number known for any bacterium (Fraser et al., 1997; Casjens et al., 2000). Among these are 12 linear and nine circular plasmids, in total 610,694 bp. Each linear plasmid of B. burgdorferi s.l. has a copy number of approximately one per chromosome (Hinnebusch and Barbour, 1992; Casjens and Huang, 1993), and many plasmids appear to contain homologous regions of DNA (Simpson et al., 1990; Barbour et al., 1996; Zuckert and Meyer, 1996; Casjens et al., 1997a). Prolonged in vitro cultivation (10–17 passages) results in the loss of some plasmids, both linear and circular. It seems likely that the lost plasmid-encoded genes are non-essential for growth in vitro. However, loss of some plasmids mediates changes in protein expression and a lost ability to infect laboratory animals (Schwan et al., 1988; Norris et al., 1995; Xu et al., 1996; Purser and Norris, 2000). Low-passage strains are therefore often more virulent than high-passage strains, and thus these plasmids may encode key virulence determinants and/or proteins involved in infection and immune evasion. This hypothesis is supported by the observation that B. burgdorferi s.l. contains plasmid-encoded genes that are selectively expressed during infection of mammalian hosts (Champion et al., 1994; Akins et al., 1995; Suk et al., 1995). The plasmids mainly contain unique genes with unknown functions, 14.5% of the plasmid genes being putative lipoprotein genes that may be involved in host–parasite interactions. Some plasmids have a high fraction of pseudogenes, possibly reflecting an ongoing rapid evolution (Casjens et al., 2000). Infectivity-associated plasmids have been identified in B. burgdorferi s.s. B31 using prolonged cultivation resulting in clones with high, low and intermediate infectivity phenotypes based on their frequency of isolation from needle-inoculated C3H/HeN mice. Thus, the plasmid lp25 had a direct correlation with infectivity as it was always present in clones of high or intermediate infectivity and was absent in all low-infectivity clones. In addition, the plasmid lp28-1, encoding the vlsE locus, also correlated with infectivity. In contrast, the plasmids cp9, cp32-3, lp21, lp28-2, lp28-4 and lp56 apparently are not required for infection in this model because clones lacking these plasmids exhibited a high-infectivity phenotype. Plasmids cp26, cp32-1, cp32-2 and/or cp32-7, cp32-4, cp32-6, cp32-8, cp32-9, lp17, lp28-3, lp36, lp38 and lp54 were consistently present in all clones examined. These results clearly indicate that the plasmids lp25 and lp28-1 encode virulence factors important for the pathogenesis of B. burgdorferi s.s. B31 (Purser and Norris, 2000).

Bacteriophages Using electron microscopic studies of B. burgdorferi s.s., the first structures resembling bacteriophages were observed and identified by Hayes et al. (1983). Earlier

66

S. Bergström et al.

studies also suggested that B. burgdorferi s.l. can carry prophages, which can be lysogenized upon treatment with ciprofloxacin (Neubert et al., 1993). However, it was not until recently that a bacteriophage designated ΦBB-1of B. burgdorferi s.l. was molecularly characterized. In that study, a DNase-protected, chloroformresistant molecule of DNA from the cell-free supernatant of a B. burgdorferi s.l. culture was recovered (Eggers and Samuels, 1999). This bacteriophage was found to be a temperate prophage derived from the 32 kb circular plasmids (cp32s) of the B. burgdorferi s.l. genome. Electron microscopy of samples from which the 32 kb DNA molecule was purified revealed bacteriophage particles (Eggers et al., 2000, 2001a,b). These authors also successfully demonstrated that ΦBB-1 could package and transduce DNA, which is the first direct evidence of a mechanism for lateral gene transfer in B. burgdorferi s.l.

Genetic tools Although sequencing of the complete genome of B. burgdorferi s.s., including all its various plasmids, has been achieved, most of the advances in pathogenesis research will come from experimental observations in the laboratory. This will not be possible unless good and reliable methods to mutagenize specific virulence-associated genes have been developed. Despite the fact that spirochaetes are important pathogens, spirochaete genetics are in their early stages and new genetic tools for studying these important pathogens are under way. Gene inactivation in spirochaetes has been achieved in B. burgdorferi s.l., Brachyspira hyodysenteriae and Treponema denticola (Tilly et al., 2000). A gene of interest can be inactivated by allelic exchange or by plasmid integration and both of these methods have been used in B. burgdorferi s.l. Random insertional mutagenesis using a transposable element has yet to be developed for spirochaetes (Berg and Berg, 1996). Until recently, genetic studies in Borrelia spp. have been hindered by the lack of an exogenous selectable marker, low transformation frequency and difficulties in growing B. burgdorferi s.l. on solid medium. The first selectable marker used for B. burgdorferi s.l. was gyrB r, a mutated form of the chromosomal gyrB gene, which encodes the B subunit of DNA gyrase and confers resistance to the antibiotic coumermycin A1 (Samuels et al., 1994a,b). The use of coumermycin as a selectable marker is limited by a high frequency of recombination with the endogenous gyrB gene. A major improvement was made when Bono et al. (2000) developed an efficient marker for mutant selection in B. burgdorferi s.l. By linking the B. burgdorferi flaA or flgB promoter with the kan gene from Tn903, transformants were resistant to high levels of kanamycin. With this innovation the pace of gene inactivation will increase, and genes that have been inactivated in B. burgdorferi s.l. and published are shown in Table 3.2. However, gene inactivation has not been achieved in an infectious background and therefore analysis of phenotypes with regard to infection has not been possible. The challenges associated with inactivating genes in infectious B. burgdorferi s.l. are not surpris-

Molecular and Cellular Biology of B. burgdorferi

67

Table 3.2. Genes that have been inactivated in B. burgdorferi. Gene

Marker

Reference

ospC oppAIV guaA rpoS gac oppAV flaB chbB

gyrB r gyrB r gyrB r gyrB r gyrB r kan gyrB r gyrB

Tilly et al. (1997) Bono et al. (1998) Tilly et al. (1998) Elias et al. (2000) Knight et al. (2000) Bono et al. (2000) Motaleb et al. (2000) Tilly et al. (2001)

ing when considering that Borrelia spp. are fastidious bacteria, with limited metabolic capacity, and infectious clones generally grow more slowly and display lower plating efficiency. Surface characteristics of the bacteria, which often correlate with infectivity, are among the obstacles to effective transformation by electroporation. Other possible reasons for reduced transformation frequency could be that a restriction/modification system is present in infectious B. burgdorferi s.l. but absent or inactive in culture-adapted bacteria. Homologous recombination may be more tightly regulated in low-passage infectious strains and recombination with a targeted locus may take place at a lower rate. Shuttle vectors have also been developed for B. burgdorferi s.l. and a broad-host-range plasmid functioning as a B. burgdorferi–E. coli shuttle vector has been reported by Sartakova et al. (2000). Stewart et al. (2001) successfully transformed infectious B. burgdorferi s.s. using a shuttle vector including a 3.3 kb region of the B. burgdorferi s.l. circular plasmid 9, and this vector could be maintained in a stable form in the spirochaete (Stewart et al., 2001). Optimal transformation frequencies are obtained with log-phase bacteria, large amounts of DNA (up to 50 µg per transformation) and high field strength (12.5–37.5 kV cm–1). Infectious B. burgdorferi isolates transform with frequencies 100-fold lower than those found for highpassage, non-infectious strains. A summary of the transformation and plating procedure is shown in Fig. 3.4.

Classification and Evolution Of the 11 known members of the B. burgdorferi s.l. species complex (Table 3.1), only three, B. burgdorferi s.s., B. afzelii and B. garinii, are known to be pathogenic in humans, but there are also unknown types of B. burgdorferi s.l. isolated from LB patients and the species defining process is still going on (Picken et al., 1996; Strle et al., 1997; Wang, G. et al., 1999a). Classification of the relapsing fever spirochaetes was based on the different argasid tick vectors involved: the ‘one tick vector species – one Borrelia species’ concept. This approach for the classification of LB spirochaetes transmitted by ixodid ticks of the I. ricinus species complex (see Eisen and Lane, Chapter 4) was abandoned when it became

68

S. Bergström et al.

10–50 µg DNA Competent Borrelia in 0.27 M sucrose, 15% glycerol 50 µl 1010 cells per ml

Incubation in 5 ml BSK H + 6% rabbit serum, 34˚ C, 24 h, without selection

Electroporation 2.5 kV, 25 µF, 200Ω

Transfer to 42oC BSKII 3% rabbit serum, 0.8% agarose + antibiotic

Bottom layer

Incubation in 34˚ C, 1% CO2 Colonies appear after 5 to 17 days Fig. 3.4. Summary of transformation and plating methodology for Borrelia burgdorferi sensu lato.

apparent that there were many exceptions (Felsenfeld, 1971; Barbour and Hayes, 1986). For example, I. ricinus transmits at least five genospecies (Hubálek and Halouzka, 1997) and B. garinii has been isolated from at least four tick species: I. ricinus (Baranton et al., 1992), Ixodes persulcatus (Kryuchechnikov et al., 1988), I. uriae (Olsen et al., 1993) and Ixodes hexagonus (Gern et al., 1997).

Molecular and Cellular Biology of B. burgdorferi

69

Each species of B. burgdorferi s.l. can be characterized to a certain extent by its vectors, host spectrum, organotropism and geographical distribution. However, speciation is mainly based on molecular methods such as phylogenetic analysis of the 5S–23S rRNA intergenic spacer sequence, 16S rRNA gene sequence, whole genome DNA reassociation experiments, analysis of rRNA gene restriction patterns, protein electrophoresis patterns and differences in reactivity to specific murine monoclonal antibodies.

Molecular characterization Numerous B. burgdorferi s.l. strains have been isolated and even more PCR fragments have been amplified from various sources. In order to determine the Borrelia species present in the samples several genotypic and phenotypic methods have been developed. Among the phenotypic methods, serotyping using monoclonal antibodies against OspA and OspC is commonly performed (Wilske et al., 1993, 1995; Wang, G. et al., 1999c). The drawback is that not all strains express all of the investigated proteins in vitro. Solely determining the protein profile on SDS-PAGE is not sufficiently reliable. Genotypic methods are not only applicable to cultivated strains but also to PCR-amplified DNA. PCR-based methods include species-specific PCR, randomly amplified polymorphic DNA (RAPD) fingerprinting or arbitrary primed (AP)-PCR, restriction fragment length polymorphism of PCR-amplified fragments (PCR-RFLP) and nucleotide sequencing (Wang, G. et al., 1999c). Methods requiring cultivation of the spirochaetes are, for example, ribotyping by hybridization with rDNA-directed probes to RFLP-generated fragments of genomic DNA, pulse field gel electrophoresis (PFGE) and the labour-intensive DNA–DNA reassociation analyses. Many methods have high congruence but differing resolution at the species and subspecies levels (Wang, G. et al., 1999c). DNA–DNA reassociation analysis is one of the reference methods for bacterial species delineation (Wayne et al., 1987). The general rule in bacterial taxonomy, that less than 70% similarity between genomes constitutes a different species, applies to B. burgdorferi s.l. (Postic et al., 1994). DNA–DNA reassociation analysis is, however, laborious to perform and therefore rarely used. Various genes have been sequenced in order to delineate and characterize B. burgdorferi s.l. In particular, the 16S rDNA, the 5S–23S rDNA intergenic spacer and the fla gene supply data representative for the whole genome (Fukunaga et al., 1996b; Wang et al., 1997). These sequences are very conserved, and in particular the 16S rRNA gene sequences in B. burgdorferi s.l are highly homologous, indicating a recent speciation (Le Fleche et al., 1997). Other gene sequences, for example p66 (Jonas Bunikis, personal communication) or the hbb gene (Valsangiacomo et al., 1997), are more variable and may contain more phylogenetic information. The outer surface protein genes ospA and ospC are less reliable in phylogeny as they are highly variable (Theisen et al., 1995; Wang et al.,

70

S. Bergström et al.

2000) and subject to lateral gene transfer (Rosa et al., 1992; Dykhuizen et al., 1993; Jauris-Heipke et al., 1995; Livey et al., 1995; Wang, G. et al., 1999b). However, it is usually possible to determine the species correctly from ospA and ospC sequences (Wang et al., 2000). Variations in the rDNA and its intergenic spacer have been used to develop different typing systems for Borrelia strains and DNA. These ribotyping methods are commonly used since they are easy to perform on both cultivated strains and B. burgdorferi s.l. DNA isolated from various sources. Analysis of restriction fragment length polymorphisms and sequencing of 16S rDNA, 23S rDNA and the intergenic spacer between the duplicated 23S and 5S rRNA genes have also been very useful in phylogenetic studies of B. burgdorferi s.l. (Marconi and Garon, 1992b; Postic et al., 1994).

Molecular epidemiology Molecular typing of B. burgdorferi s.l. DNA and strains from ticks and animals are important tools in epidemiology. For example, we have studied the importance of seabirds for transport of B. burgdorferi s.l. over long distances, and using molecular methods have identified B. burgdorferi s.l. DNA in seabird ticks, I. uriae (Olsen et al., 1995). The seabird tick, I. uriae, has a circumpolar distribution in both the northern and southern hemispheres and in this study identical B. garinii flagellin gene (flaB) sequences were detected in I. uriae from these hemispheres, indicating a transequatorial transport of B. garinii. Phylogenetic analysis of I. uriae ITS2 and 16S rDNA sequences suggested that northern and southern I. uriae might be reproductively separated. Therefore, passive transport of infected ticks between the polar regions is unlikely and instead seabirds probably carry an active B. burgdorferi s.l. infection during their migration. Furthermore, in another study we have demonstrated that B. burgdorferi s.l. can persist for several months in passerine birds and the infection in redwing thrushes can be reactivated in response to migration (Gylfe et al., 2000). Thus, birds may be more infectious to ticks during their migration and therefore important long-range disseminators of B. burgdorferi s.l. This work shows that migrating seabirds and passerine birds are probably important for the long-range dispersal of B. burgdorferi s.l. and that this mechanism of dispersal could be important for the distribution of human Lyme disease. Using similar approaches, other researchers have demonstrated the relative importance of different animals as reservoir hosts for different B. burgdorferi s.l species in Europe. For example, since Borrelia valaisiana has only been isolated from birds, birds have been implicated as being the only reservoir for this Borrelia species (Kurtenbach et al., 1998). B. afzelii, on the other hand, is more frequently isolated from small rodents and ticks in areas where rodents are abundant, suggesting that rodents are the main reservoir hosts for B. afzelii (Hu et al., 1997). In areas where birds are important as tick hosts, B. garinii and B. valaisiana predominate (Kurtenbach et al., 1998; Gray et al., 2000). There seem to be at least

Molecular and Cellular Biology of B. burgdorferi

71

two distinct groups of B. garinii: the eastern type, which is mainly infectious to mice, and the western type, which is mainly infectious to birds (see Kurtenbach et al., Chapter 5).

Evolution of B. burgdorferi s.l. Since the 1980s, bacterial genomics have been performed in order to investigate the evolution of the members of the bacterial domain. Using very conserved genes, it has also been possible to compare the evolution between all the three domains: Bacteria, Archae and Eukarya. Spirochaetes have been separated into eight different genera based on habitat, pathogenicity, rRNA sequences and morphological and physiological characters. They belong to a very disparate group of microorganisms that is found in a variety of ecological niches such as soil, marine sediments, in arthropods or as parasites of vertebrates. The spirochaetes constitute a major bacterial grouping that by multicellular eukaryotic standards would be considered at least a division or phylum (Woese, 1987; Olsen et al., 1994). In a recent review by Carl Woese et al. (2000), the aminoacyl-tRNA synthetases were used very elegantly to construct a picture of the evolutionary process in bacteria. This was done to obtain a clearer concept of how modern cells have evolved and to identify various stages in the process of bacterial evolution (Woese et al., 2000). The genome sequence of two pathogenic spirochaetes, B. burgdorferi s.s. and T. pallidum, was recently published. The sequenced genomes from these two members of the spirochaetal division do not cover the full phylogenetic extent of this division but they are well separated phylogenetically. These two spirochaetes contain 19 of the aminoacyl-tRNA synthetases (they have no GlnRS), and from the comparison of these genes it was evident that in only two cases, threonine and proline, are the spirochaete enzymes clearly evolutionarily separated from one another. Interestingly, it was also apparent from this analysis that the spirochaetes have seven out of 19 aminoacyl-tRNA synthetases that are of both archaeal and eukaryotic origin. Although the treponemes and borreliae exhibit a lot of similarities, a striking difference is their genome organization and structure. The members of the genus Borrelia have a very unusual genome organization among prokaryotic organisms as they have a polyploid genome that is mostly linear. The small linear replicons in borreliae have been called plasmids but, as in the genus Trypanosoma, there is justification for designating them mini chromosomes instead. A comparison of the ends of these replicons, i.e. the telomeres, with telomeric sequences of other linear double-stranded DNA replicons revealed sequence similarities with poxviruses, especially with the iridovirus agent of African swine fever. Interestingly, the latter virus and a relapsing fever Borrelia sp., Borrelia duttoni, share the same tick vector. These findings suggest that the linear replicons of B. burgdorferi s.l. originated through a horizontal gene transfer across the kingdom. Since this structural feature is found throughout the genus Borrelia, the evolutionary change must have occurred a very long time ago (Hinnebusch and Barbour, 1992).

72

S. Bergström et al.

An interesting hypothesis on the evolution of Borrelia spp. spirochaetes is that they have co-evolved together with their different arthropod vectors. The concept of the ‘one vector – one Borrelia species’ has been accepted in relapsing fever Borrelia spp. as its original species designation was related to its specific tick vector. Despite the fact that, as mentioned previously, this concept cannot be applied to LB borreliae, several phylogenetic analyses of B. burgdorferi s.l. and its different ixodid tick vectors have indicated that these organisms might have co-evolved. In a study by Fukunaga et al. (2000), an internal transcribed spacer (ITS2) sequence between the 5.8S and 28S rRNA genes was used to determine the phylogenetic relationships among Ixodes spp. vectors of LB-causing Borrelia spp. spirochaetes. From these analyses it was concluded that the Borrelia spp. associated with LB are found mainly in ticks of the I. ricinus species complex, whereas other closely related tick species, e.g. I. uriae, not known to transmit the Borrelia spp. that cause LB in humans, appear to have a specific association with other closely related Borrelia spp. Thus, from these studies it was found that there is a large similarity in the phylogeny of the Borrelia taxa and the phylogenetic relationships among Ixodes spp. ticks. The pattern of distribution of LB is comparable with a multifocal worldwide epidemic with the endemic regions extending over the continents. However, most species of the B. burgdorferi s.l. complex have a relatively limited distribution, the only exception being B. burgdorferi s.s., which is found both in Europe and in the USA. The question is, from where and how have the LB Borrelia spp. migrated? The larger genetic variability and the clinical polymorphism seen in the Old World suggest that the B. burgdorferi s.l. complex emanated from this geographical area. However, in a series of experiments Baranton and co-workers compared the genetic polymorphism of B. burgdorferi s.s. strains isolated in Europe or the USA, using several molecular methods (Foretz et al., 1997; Marti Ras et al., 1997). A genetic distance method and a parsimony method were used to perform phylogenetic analysis of each of the two sets of data. The results were consistent and showed that the American B. burgdorferi s.s. strains were more heterogeneous than the European ones; thus, it might be possible that this species evolved in America and was introduced to Europe, and that the different populations have evolved separately. In contrast to the distribution of B. burgdorferi s.s., B. afzelii appears to be more localized and stable within regions in Europe and Asia where it does not spread over large distances. This is probably due to its vector and host relationships. B. garinii, on the other hand, was found to be spread over large geographical regions in Europe and was associated with both the ticks within the I. ricinus complex and the seabird tick I. uriae involved in the maintenance of a novel, seabirdrelated, enzootic cycle of LB (Olsen et al., 1993). In an attempt to delineate the expansion and molecular polymorphism of B. burgdorferi s.l. in the north of Europe, Borrelia strains recovered from different tick species on islands in the northern Baltic Sea and North Atlantic were genotypically and phenotypically characterized. Molecular polymorphism of B. garinii was shown to be a result of interactions among several regional enzootic foci of this genospecies.

Molecular and Cellular Biology of B. burgdorferi

73

Concluding Comments The aim of this chapter has been to describe the current status of the molecular and cellular biology of LB spirochaetes. Much of this recent knowledge is directly coupled to the published genome sequence of B. burgdorferi s.s. B31. The new genomics and proteomics era that we have entered will be instrumental for future understanding of B. burgdorferi s.l. In addition, the combination of data from other microorganisms also has the potential of further increasing knowledge of the biology of Borrelia spp. spirochaetes in general and their pathogenesis in particular. The approach in the future will be to identify the central regulators of common host responses and the manner in which Borrelia spp. spirochaetes attempt to defeat this process. This should lead to the development of valuable tools for designing novel prophylactic and diagnostic methods.

References Adam, T., Gassmann, G.S., Rasiah, C. and Gobel, U.B. (1991) Phenotypic and genotypic analysis of Borrelia burgdorferi isolates from various sources. Infection and Immunity 59, 2579–2585. Akins, D.R., Porcella, S.F., Popova, T.G., Shevchenko, D., Baker, S.I., Li, M., Norgard, M.V. and Radolf, J.D. (1995) Evidence for in vivo but not in vitro expression of a Borrelia burgdorferi outer surface protein F (OspF) homologue. Molecular Microbiology 18, 507–520. Akins, D.R., Bourell, K.W., Caimano, M.J., Norgard, M.V. and Radolf, J.D. (1998) A new animal model for studying Lyme disease spirochetes in a mammalian host-adapted state. Journal of Clinical Investigation 101, 2240–2250. Akins, D.R., Caimano, M.J., Yang, X., Cerna, F., Norgard, M.V. and Radolf, J.D. (1999) Molecular and evolutionary analysis of Borrelia burgdorferi 297 circular plasmid-encoded lipoproteins with OspE- and OspF-like leader peptides. Infection and Immunity 67, 1526–1532. Alban, P.S., Johnson, P.W. and Nelson, D.R. (2000) Serum-starvation-induced changes in protein synthesis and morphology of Borrelia burgdorferi. Microbiology 146, 119–127.

Allardet-Servent, A., Michaux-Charachon, S., Jumas-Bilak, E., Karayan, L. and Ramuz, M. (1993) Presence of one linear and one circular chromosome in the Agrobacterium tumefaciens C58 genome. Journal of Bacteriology 175, 7869–7874. Anderson, J.F., Magnarelli, L.A., LeFebvre, R.B., Andreadis, T.G., McAninch, J.B., Perng, G.C. and Johnson, R.C. (1989) Antigenically variable Borrelia burgdorferi isolated from cottontail rabbits and Ixodes dentatus in rural and urban areas. Journal of Clinical Microbiology 27, 13–20. Anguita, J., Thomas, V., Samanta, S., Persinski, R., Hernanz, C., Barthold, S.W. and Fikrig, E. (2001) Borrelia burgdorferi-induced inflammation facilitates spirochete adaptation and variable major protein-like sequence locus recombination. Journal of Immunology 167, 3383–3390. Baranton, G., Postic, D., Saint Girons, I., Boerlin, P., Piffaretti, J.C., Assous, M. and Grimont, P.A. (1992) Delineation of Borrelia burgdorferi sensu stricto, Borrelia garinii sp. nov., and group VS461 associated with Lyme borreliosis. International Journal of Systematic Bacteriology 42, 378–383. Baranton, G., Marti Ras, N. and Postic, D.

74

S. Bergström et al.

(1998) Borrelia burgdorferi, taxonomy, pathogenicity and spread. Annales de Médecine Interne (Paris) 149, 455–458 (in French). Barbour, A.G. (1988) Plasmid analysis of Borrelia burgdorferi, the Lyme disease agent. Journal of Clinical Microbiology 26, 475–478. Barbour, A.G. (1990) Antigenic variation of a relapsing fever Borrelia species. Annual Reviews in Microbiology 44, 155–171. Barbour, A.G. (1993) Linear DNA of Borrelia species and antigenic variation. Trends in Microbiology 1, 236–239. Barbour, A.G. and Garon, C.F. (1987) Linear plasmids of the bacterium Borrelia burgdorferi have covalently closed ends. Science 237, 409–411. Barbour, A.G. and Garon, C.F. (1988) The genes encoding major surface proteins of Borrelia burgdorferi are located on a plasmid. Annals of the New York Academy of Sciences 539, 144–153. Barbour, A.G. and Hayes, S.F. (1986) Biology of Borrelia species. Microbiological Reviews 50, 381–400. Barbour, A.G., Tessier, S.L. and Hayes, S.F. (1984) Variation in a major surface protein of Lyme disease spirochetes. Infection and Immunity 45, 94–100. Barbour, A.G., Hayes, S.F., Heiland, R.A., Schrumpf, M.E. and Tessier, S.L. (1986) A Borrelia-specific monoclonal antibody binds to a flagellar epitope. Infection and Immunity 52, 549–554. Barbour, A.G., Burman, N., Carter, C.J., Kitten, T. and Bergstrom, S. (1991a) Variable antigen genes of the relapsing fever agent Borrelia hermsii are activated by promoter addition. Molecular Microbiology 5, 489–493. Barbour, A.G., Carter, C.J., Burman, N., Freitag, C.S., Garon, C.F. and Bergstrom, S. (1991b) Tandem insertion sequence-like elements define the expression site for variable antigen genes of Borrelia hermsii. Infection and Immunity 59, 390–397. Barbour, A.G., Carter, C.J., Bundoc, V.

and Hinnebusch, J. (1996) The nucleotide sequence of a linear plasmid of Borrelia burgdorferi reveals similarities to those of circular plasmids of other prokaryotes. Journal of Bacteriology 178, 6635–6639. Baril, C., Richaud, C., Baranton, G. and Saint Girons, I.S. (1989) Linear chromosome of Borrelia burgdorferi. Research in Microbiology 140, 507–516. Beck, G., Habicht, G.S., Benach, J.L. and Coleman, J.L. (1985) Chemical and biologic characterization of a lipopolysaccharide extracted from the Lyme disease spirochete (Borrelia burgdorferi). Journal of Infectious Diseases 152, 108–117. Belisle, J.T., Brandt, M.E., Radolf, J.D. and Norgard, M.V. (1994) Fatty acids of Treponema pallidum and Borrelia burgdorferi lipoproteins. Journal of Bacteriology 176, 2151–2157. Berg, C.M. and Berg, D.E. (1996) Transposable element tools for microbial genetics. In: Neidhardt, F.C., Curtis, R.C. III, Ingraham, J.L., Lin, E.C.C., Low, K.B. et al. (eds) Escherichia coli and Salmonella. ASM Press, Washington, DC, pp. 2588–2612. Berg, S. (2001) The glucolipid pathway in Acholeplasma laidlawii. Functional and structural properties of three consecutively-acting enzymes and similarities to orthologs in prokaryotes. PhD thesis, Umeå University, Umeå, Sweden. ISBN 91-7305-025-3. Bergström, S., Bundoc, V.G. and Barbour, A.G. (1989) Molecular analysis of linear plasmid-encoded major surface proteins, OspA and OspB, of the Lyme disease spirochaete Borrelia burgdorferi. Molecular Microbiology 3, 479–486. Bergström, S., Garon, C.F., Barbour, A.G. and MacDougall, J. (1992) Extrachromosomal elements of spirochetes. Research in Microbiology 143, 623–628. Bessler, W.G., Cox, M., Lex, A., Suhr, B., Wiesmuller, K.H. and Jung, G. (1985) Synthetic lipopeptide analogs of

Molecular and Cellular Biology of B. burgdorferi

bacterial lipoprotein are potent polyclonal activators for murine B lymphocytes. Journal of Immunology 135, 1900–1905. Bledsoe, H.A., Carroll, J.A., Whelchel, T.R., Farmer, M.A., Dorward, D.W. and Gherardini, F.C. (1994) Isolation and partial characterization of Borrelia burgdorferi inner and outer membranes by using isopycnic centrifugation. Journal of Bacteriology 176, 7447–7455. Bono, J.L., Tilly, K., Stevenson, B., Hogan, D. and Rosa, P. (1998) Oligopeptide permease in Borrelia burgdorferi: putative peptide-binding components encoded by both chromosomal and plasmid loci. Microbiology 144, 1033–1044. Bono, J.L., Elias, A.F., Kupko, J.J. III, Stevenson, B., Tilly, K. and Rosa, P. (2000) Efficient targeted mutagenesis in Borrelia burgdorferi. Journal of Bacteriology 182, 2445–2452. Borst, P. (1991) Molecular genetics of antigenic variation. Immunology Today 12, A29–A33. Brahamsha, B. and Greenberg, E.P. (1988) Biochemical and cytological analysis of the complex periplasmic flagella from Spirochaeta aurantia. Journal of Bacteriology 170, 4023–4032. Brandt, M.E., Riley, B.S., Radolf, J.D. and Norgard, M.V. (1990) Immunogenic integral membrane proteins of Borrelia burgdorferi are lipoproteins. Infection and Immunity 58, 983–991. Brown, E.L., Wooten, R.M., Johnson, B.J., Iozzo, R.V., Smith, A., Dolan, M.C., Guo, B.P., Weis, J.J. and Hook, M. (2001) Resistance to Lyme disease in decorin-deficient mice. Journal of Clinical Investigation 107, 845–852. Brusca, J.S., McDowall, A.W., Norgard, M.V. and Radolf, J.D. (1991) Localization of outer surface proteins A and B in both the outer membrane and intracellular compartments of Borrelia burgdorferi. Journal of Bacteriology 173, 8004–8008. Bunikis, J., Noppa, L. and Bergström, S.

75

(1995) Molecular analysis of a 66-kDa protein associated with the outer membrane of Lyme disease Borrelia. FEMS Microbiology Letters 131, 139–145. Burgdorfer, W., Barbour, A.G., Hayes, S.F., Benach, J.L., Grunwaldt, E. and Davis, J.P. (1982) Lyme disease – a tickborne spirochetosis? Science 216, 1317–1319. Burman, N., Bergström, S., Restrepo, B.I. and Barbour, A.G. (1990) The variable antigens Vmp7 and Vmp21 of the relapsing fever bacterium Borrelia hermsii are structurally analogous to the VSG proteins of the African trypanosome. Molecular Microbiology 4, 1715–1726. Canica, M.M., Nato, F., du Merle, L., Mazie, J.C., Baranton, G. and Postic, D. (1993) Monoclonal antibodies for identification of Borrelia afzelii sp. nov. associated with late cutaneous manifestations of Lyme borreliosis. Scandinavian Journal of Infectious Diseases 25, 441–448. Carroll, J.A. and Gherardini, F.C. (1996) Membrane protein variations associated with in vitro passage of Borrelia burgdorferi. Infection and Immunity 64, 392–398. Carroll, J.A., El-Hage, N., Miller, J.C., Babb, K. and Stevenson, B. (2001) Borrelia burgdorferi RevA antigen is a surface-exposed outer membrane protein whose expression is regulated in response to environmental temperature and pH. Infection and Immunity 69, 5286–5293. Carter, C.J., Bergstrom, S., Norris, S.J. and Barbour, A.G. (1994) A family of surface-exposed proteins of 20 kilodaltons in the genus Borrelia. Infection and Immunity 62, 2792–2799. Casjens, S. and Huang, W.M. (1993) Linear chromosomal physical and genetic map of Borrelia burgdorferi, the Lyme disease agent. Molecular Microbiology 8, 967–980. Casjens, S., van Vugt, R., Tilly, K., Rosa, P.A. and Stevenson, B. (1997a) Homology throughout the multiple 32kilobase circular plasmids present in

76

S. Bergström et al.

Lyme disease spirochetes. Journal of Bacteriology 179, 217–227. Casjens, S., Murphy, M., DeLange, M., Sampson, L., van Vugt, R. and Huang, W.M. (1997b) Telomeres of the linear chromosomes of Lyme disease spirochaetes: nucleotide sequence and possible exchange with linear plasmid telomeres. Molecular Microbiology 26, 581–596. Casjens, S., Palmer, N., van Vugt, R., Huang, W.M., Stevenson, B., Rosa, P., Lathigra, R., Sutton, G., Peterson, J., Dodson, R.J., Haft, D., Hickey, E., Gwinn, M., White, O. and Fraser, C.M. (2000) A bacterial genome in flux: the twelve linear and nine circular extrachromosomal DNAs in an infectious isolate of the Lyme disease spirochete Borrelia burgdorferi. Molecular Microbiology 35, 490–516. Chaconas, G., Stewart, P.E., Tilly, K., Bono, J.L. and Rosa, P. (2001) Telomere resolution in the Lyme disease spirochete. EMBO Journal 20, 3229–3237. Chamberlain, N.R., DeOgny, L., Slaughter, C., Radolf, J.D. and Norgard, M.V. (1989) Acylation of the 47-kilodalton major membrane immunogen of Treponema pallidum determines its hydrophobicity. Infection and Immunity 57, 2878–2885. Champion, C.I., Blanco, D.R., Skare, J.T., Haake, D.A., Giladi, M., Foley, D., Miller, J.N. and Lovett, M.A. (1994) A 9.0-kilobase-pair circular plasmid of Borrelia burgdorferi encodes an exported protein: evidence for expression only during infection. Infection and Immunity 62, 2653–2661. Charon, N.W., Greenberg, E.P., Koopman, M.B. and Limberger, R.J. (1992) Spirochete chemotaxis, motility, and the structure of the spirochetal periplasmic flagella. Research in Microbiology 143, 597–603. Chen, J., Field, J.A., Glickstein, L., Molloy, P.J., Huber, B.T. and Steere, A.C.

(1999) Association of antibiotic treatment-resistant Lyme arthritis with T cell responses to dominant epitopes of outer surface protein A of Borrelia burgdorferi. Arthritis and Rheumatism 42, 1813–1822. Cinco, M., Banfi, E., Balanzin, D., Godeas, C. and Panfili, E. (1991) Evidence for (lipo) oligosaccharides in Borrelia burgdorferi and their serological specificity. FEMS Immunology and Medical Microbiology 3, 33–38. Cockayne, A., Bailey, M.J. and Penn, C.W. (1987) Analysis of sheath and core structures of the axial filament of Treponema pallidum. Journal of General Microbiology 133, 1397–1407. Comstock, L.E., Fikrig, E., Shoberg, R.J., Flavell, R.A. and Thomas, D.D. (1993) A monoclonal antibody to OspA inhibits association of Borrelia burgdorferi with human endothelial cells. Infection and Immunity 61, 423–431. Cowan, S.W., Schirmer, T., Rummel, G., Steiert, M., Ghosh, R., Pauptit, R.A., Jansonius, J.N. and Rosenbusch, J.P. (1992) Crystal structures explain functional properties of two E. coli porins. Nature 358, 727–733. Cox, D.L., Chang, P., McDowall, A.W. and Radolf, J.D. (1992) The outer membrane, not a coat of host proteins, limits antigenicity of virulent Treponema pallidum. Infection and Immunity 60, 1076–1083. Cox, D.L., Akins, D.R., Bourell, K.W., Lahdenne, P., Norgard, M.V. and Radolf, J.D. (1996) Limited surface exposure of Borrelia burgdorferi outer surface lipoproteins. Proceedings of the National Academy of Sciences, USA 93, 7973–7978. Crespi, M., Messens, E., Caplan, A.B., van Montagu, M. and Desomer, J. (1992) Fasciation induction by the phytopathogen Rhodococcus fascians depends upon a linear plasmid encoding a cytokinin synthase gene. EMBO Journal 11, 795–804. Das, S., Shraga, D., Gannon, C., Lam,

Molecular and Cellular Biology of B. burgdorferi

T.T., Feng, S., Brunet, L.R., Telford, S.R., Barthold, S.W., Flavell, R.A. and Fikrig, E. (1996) Characterization of a 30-kDa Borrelia burgdorferi substratebinding protein homologue. Research in Microbiology 147, 739–751. Davidson, B.E., MacDougall, J. and Saint Girons, I. (1992) Physical map of the linear chromosome of the bacterium Borrelia burgdorferi 212, a causative agent of Lyme disease, and localization of rRNA genes. Journal of Bacteriology 174, 3766–3774. Deres, K., Schild, H., Wiesmuller, K.H., Jung, G. and Rammensee, H.G. (1989) In vivo priming of virus-specific cytotoxic T lymphocytes with synthetic lipopeptide vaccine. Nature 342, 561–564. Dorward, D.W., Schwan, T.G. and Garon, C.F. (1991) Immune capture and detection of Borrelia burgdorferi antigens in urine, blood, or tissues from infected ticks, mice, dogs, and humans. Journal of Clinical Microbiology 29, 1162–1170. Dunn, J.J., Lade, B.N. and Barbour, A.G. (1990) Outer surface protein A (OspA) from the Lyme disease spirochete, Borrelia burgdorferi: high level expression and purification of a soluble recombinant form of OspA. Protein Expression and Purification 1, 159–168. Duray, P.H. (1992) Target organs of B. burgdorferi infections: functional responses and histology. In: Schutzer, S.E. (ed.) Lyme Disease: Molecular and Immunologic Approaches. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York, pp. 11–30. Dykhuizen, D.E., Polin, D.S., Dunn, J.J., Wilske, B., Preac-Mursic, V., Dattwyler, R.J. and Luft, B.J. (1993) Borrelia burgdorferi is clonal: implications for taxonomy and vaccine development. Proceedings of the National Academy of Sciences, USA 90, 10163–10167. Eggers, C.H. and Samuels, D.S. (1999) Molecular evidence for a new bacteriophage of Borrelia burgdorferi. Journal of Bacteriology 181, 7308–7313.

77

Eggers, C.H., Casjens, S., Hayes, S.F., Garon, C.F., Damman, C.J., Oliver, D.B. and Samuels, D.S. (2000) Bacteriophages of spirochetes. Journal of Molecular Microbiology and Biotechnology 2, 365–373. Eggers, C.H., Casjens, S. and Samuels, D.S. (2001a) Bacteriophages of B. burgdorferi and other spirochetes. In: Saier, M.H. Jr and García-Lara, J. (eds) The Spirochetes: Molecular and Cellular Biology. Horizon Press, Wymondham, UK, p. 222. Eggers, C.H., Kimmel, B.J., Bono, J.L., Elias, A.F., Rosa, P. and Samuels, D.S. (2001b) Transduction by phiBB-1, a bacteriophage of Borrelia burgdorferi. Journal of Bacteriology 183, 4771–4778. Eiffert, H., Lotter, H., Jarecki-Khan, K. and Thomssen, R. (1991) Identification of an immunoreactive non-proteinaceous component in Borrelia burgdorferi. Medical Microbiology and Immunology 180, 229–237. El-Hage, N., Lieto, L.D. and Stevenson, B. (1999) Stability of erp loci during Borrelia burgdorferi infection: recombination is not required for chronic infection of immunocompetent mice. Infection and Immunity 67, 3146–3150. Elias, A.F., Bono, J.L., Carroll, J.A., Stewart, P., Tilly, K. and Rosa, P. (2000) Altered stationary-phase response in a Borrelia burgdorferi rpoS mutant. Journal of Bacteriology 182(10), 2909–2918. Erdile, L.F., Brandt, M.A., Warakomski, D.J., Westrack, G.J., Sadziene, A., Barbour, A.G. and Mays, J.P. (1993) Role of attached lipid in immunogenicity of Borrelia burgdorferi OspA. Infection and Immunity 61, 81–90. Felsenfeld, O. (1971) Borrelia: Strains, Vectors, Human and Animal Borreliosis. Warren H. Green Inc., St Louis, Missouri, 180 pp. Feng, S., Hodzic, E., Stevenson, B. and Barthold, S.W. (1998) Humoral immunity to Borrelia burgdorferi N40 decorin

78

S. Bergström et al.

binding proteins during infection of laboratory mice. Infection and Immunity 66, 2827–2835. Ferdows, M.S. and Barbour, A.G. (1989) Megabase-sized linear DNA in the bacterium Borrelia burgdorferi, the Lyme disease agent. Proceedings of the National Academy of Sciences, USA 86, 5969–5973. Finlay, B.B. and Falkow, S. (1989) Common themes in microbial pathogenicity. Microbiology Reviews 53, 210–230. Foretz, M., Postic, D. and Baranton, G. (1997) Phylogenetic analysis of Borrelia burgdorferi sensu stricto by arbitrarily primed PCR and pulsed-field gel electrophoresis. International Journal of Systematic Bacteriology 47, 11–18. Fraser, C.M., Gocayne, J.D., White, O., Adams, M.D., Clayton, R.A., Fleischmann, R.D., Bult, C.J., Kerlavage, A.R., Sutton, G., Kelley, J.M. et al. (1995) The minimal gene complement of Mycoplasma genitalium. Science 270, 397–403. Fraser, C.M., Casjens, S., Huang, W.M., Sutton, G.G., Clayton, R., Lathigra, R., White, O., Ketchum, K.A., Dodson, R., Hickey, E.K., Gwinn, M., Dougherty, B., Tomb, J.F., Fleischmann, R.D., Richardson, D., Peterson, J., Kerlavage, A.R., Quackenbush, J., Salzberg, S., Hanson, M., van Vugt, R., Palmer, N., Adams, M.D., Gocayne, J., Weidman, J., Utterback, T., Watthey, L., McDonald, L., Artiach, P., Bowman, C., Garland, S., Fujii, C., Cotton, M.D., Horst, K., Roberts, K., Hatch, B., Smith, H.O. and Venter, J.C. (1997) Genomic sequence of a Lyme disease spirochaete, Borrelia burgdorferi. Nature 390, 580–586. Fraser, C.M., Norris, S.J., Weinstock, G.M., White, O., Sutton, G.G., Dodson, R., Gwinn, M., Hickey, E.K., Clayton, R., Ketchum, K.A., Sodergren, E., Hardham, J.M., McLeod, M.P., Salzberg, S., Peterson, J., Khalak, H., Richardson, D., Howell,

J.K., Chidambaram, M., Utterback, T., McDonald, L., Artiach, P., Bowman, C., Cotton, M., Fujii, C., Garland, S., Hatch, B., Horst, K., Roberts, K., Sandusky, M., Weidman, J., Smith, H.O. and Venter, J.C. (1998) Complete genome sequence of Treponema pallidum, the syphilis spirochete. Science 281, 375–388. Fukunaga, M., Hamase, A., Okada, K. and Nakao, M. (1996a) Borrelia tanukii sp. nov. and Borrelia turdae sp. nov. found from ixodid ticks in Japan: rapid species identification by 16S rRNA gene-targeted PCR analysis. Microbiology and Immunology 40, 877–881. Fukunaga, M., Okada, K., Nakao, M., Konishi, T. and Sato, Y. (1996b) Phylogenetic analysis of Borrelia species based on flagellin gene sequences and its application for molecular typing of Lyme disease borreliae. International Journal of Systematic Bacteriology 46, 898–905. Fukunaga, M., Yabuki, M., Hamase, A., Oliver, J.H. Jr and Nakao, M. (2000) Molecular phylogenetic analysis of ixodid ticks based on the ribosomal DNA spacer, internal transcribed spacer 2, sequences. Journal of Parasitology 86, 38–43. Garcia-Monco, J.C., Fernandez-Villar, B. and Benach, J.L. (1989) Adherence of the Lyme disease spirochete to glial cells and cells of glial origin. Journal of Infectious Diseases 160, 497–506. Garon, C.F., Dorward, D.W. and Corwin, M.D. (1989) Structural features of Borrelia burgdorferi – the Lyme disease spirochete: silver staining for nucleic acids. Scanning Microscopy 3 (suppl.), 109–115. Ge, Y. and Charon, N.W. (1997a) An unexpected flaA homolog is present and expressed in Borrelia burgdorferi. Journal of Bacteriology 179, 552–556. Ge, Y. and Charon, N.W. (1997b) Molecular characterization of a flagellar/chemotaxis operon in the spirochete

Molecular and Cellular Biology of B. burgdorferi

Borrelia burgdorferi. FEMS Microbiology Letters 153, 425–431. Ge, Y., Old, I., Saint Girons, I., Yelton, D.B. and Charon, N.W. (1996) FliH and fliI of Borrelia burgdorferi are similar to flagellar and virulence factor export proteins of other bacteria. Gene 168, 73–75. Ge, Y., Old, I.G., Girons, I.S. and Charon, N.W. (1997) The flgK motility operon of Borrelia burgdorferi is initiated by a sigma 70-like promoter. Microbiology 143, 1681–1690. Ge, Y., Li, C., Corum, L., Slaughter, C.A. and Charon, N.W. (1998) Structure and expression of the FlaA periplasmic flagellar protein of Borrelia burgdorferi. Journal of Bacteriology 180, 2418–2425. Gern, L., Rouvinez, E., Toutoungi, L.N. and Godfroid, E. (1997) Transmission cycles of Borrelia burgdorferi sensu lato involving Ixodes ricinus and/or I. hexagonus ticks and the European hedgehog, Erinaceus europaeus, in suburban and urban areas in Switzerland. Folia Parasitologica 44, 309–314. Gilmore, R.D. and Piesman, J. (2000) Inhibition of Borrelia burgdorferi migration from the midgut to the salivary glands following feeding by ticks on OspCimmunized mice. Infection and Immunity 68, 411–414. Gilmore, R.D. Jr, Kappel, K.J., Dolan, M.C., Burkot, T.R. and Johnson, B.J. (1996) Outer surface protein C (OspC), but not P39, is a protective immunogen against a tick-transmitted Borrelia burgdorferi challenge: evidence for a conformational protective epitope in OspC. Infection and Immunity 64, 2234–2239. Goldstein, S.F., Charon, N.W. and Kreiling, J.A. (1994) Borrelia burgdorferi swims with a planar waveform similar to that of eukaryotic flagella. Proceedings of the National Academy of Sciences, USA 91, 3433–3437. Goldstein, S.F., Buttle, K.F. and Charon, N.W. (1996) Structural analysis of the Leptospiraceae and Borrelia burgdorferi by

79

high-voltage electron microscopy. Journal of Bacteriology 178, 6539–6545. Gominet, M., Slamti, L., Gilois, N., Rose, M. and Lereclus, D. (2001) Oligopeptide permease is required for expression of the Bacillus thuringiensis plcR regulon and for virulence. Molecular Microbiology 40, 963–975. Goodell, E.W. and Higgins, C.F. (1987) Uptake of cell wall peptides by Salmonella typhimurium and Escherichia coli. Journal of Bacteriology 169, 3861–3865. Gorbacheva, V.Y., Godfrey, H.P. and Cabello, F.C. (2000) Analysis of the bmp gene family in Borrelia burgdorferi sensu lato. Journal of Bacteriology 182, 2037–2042. Gray, J.S., Robertson, J.N. and Key, S. (2000) Limited role of rodents as reservoirs of Borrelia burgdorferi sensu lato in Ireland. European Journal of Epidemiology 16, 101–103. Gross, D.M., Forsthuber, T., TaryLehmann, M., Etling, C., Ito, K., Nagy, Z.A., Field, J.A., Steere, A.C. and Huber, B.T. (1998) Identification of LFA-1 as a candidate autoantigen in treatment-resistant Lyme arthritis. Science 281, 703–706. Guo, B.P., Norris, S.J., Rosenberg, L.C. and Hook, M. (1995) Adherence of Borrelia burgdorferi to the proteoglycan decorin. Infection and Immunity 63, 3467–3472. Guo, B.P., Brown, E.L., Dorward, D.W., Rosenberg, L.C. and Hook, M. (1998) Decorin-binding adhesins from Borrelia burgdorferi. Molecular Microbiology 30, 711–723. Gylfe, Å., Olsen, B., Strasevicius, D., Marti Ras, N., Weihe, P., Noppa, L., Östberg, Y., Baranton, G. and Bergström, S. (1999) Isolation of Lyme disease Borrelia from puffins (Fratercula arctica) and seabird ticks (Ixodes uriae) on the Faeroe islands. Journal of Clinical Microbiology 37, 890–896. Gylfe, Å., Bergström, S., Lundström, J. and Olsen, B. (2000) Reactivation of

80

S. Bergström et al.

Borrelia infection in birds. Nature 403, 724–725. Haas, R. and Meyer, T.F. (1986) The repertoire of silent pilus genes in Neisseria gonorrhoeae: evidence for gene conversion. Cell 44, 107–115. Hagblom, P., Segal, E., Billyard, E. and So, M. (1985) Intragenic recombination leads to pilus antigenic variation in Neisseria gonorrhoeae. Nature 315, 156–158. Hagman, K.E., Lahdenne, P., Popova, T.G., Porcella, S.F., Akins, D.R., Radolf, J.D. and Norgard, M.V. (1998) Decorin-binding protein of Borrelia burgdorferi is encoded within a two-gene operon and is protective in the murine model of Lyme borreliosis. Infection and Immunity 66, 2674–2683. Hanson, M.S., Cassatt, D.R., Guo, B.P., Patel, N.K., McCarthy, M.P., Dorward, D.W. and Hook, M. (1998) Active and passive immunity against Borrelia burgdorferi decorin binding protein A (DbpA) protects against infection. Infection and Immunity 66, 2143–2153. Hayashi, S. and Wu, H.C. (1990) Lipoproteins in bacteria. Journal of Bioenergetics and Biomembranes 22, 451–471. Hayes, S.F., Burgdorfer, W. and Barbour, A.G. (1983) Bacteriophage in the Ixodes dammini spirochete, etiological agent of Lyme disease. Journal of Bacteriology 154, 1436–1439. Hellwage, J., Meri, T., Heikkila, T., Alitalo, A., Panelius, J., Lahdenne, P., Seppala, I.J. and Meri, S. (2001) The complement regulator factor H binds to the surface protein OspE of Borrelia burgdorferi. Journal of Biological Chemistry 276, 8427–8435. Hinnebusch, J. and Barbour, A.G. (1992) Linear- and circular-plasmid copy numbers in Borrelia burgdorferi. Journal of Bacteriology 174, 5251–5257. Hinnebusch, J., Bergstrom, S. and Barbour, A.G. (1990) Cloning and sequence analysis of linear plasmid telomeres of the bacterium Borrelia burgdorferi. Molecular Microbiology 4, 811–820.

Hoffmann, P., Heinle, S., Schade, U.F., Loppnow, H., Ulmer, A.J., Flad, H.D., Jung, G. and Bessler, W.G. (1988) Stimulation of human and murine adherent cells by bacterial lipoprotein and synthetic lipopeptide analogues. Immunobiology 177, 158–170. Holt, S.C. (1978) Anatomy and chemistry of spirochetes. Microbiological Reviews 42, 114–160. Hossain, H., Wellensiek, H., Geyer, R. and Lochnit, G. (2001) Structural analysis of glycolipids from Borrelia burgdorferi. Biochimie 83, 683–692. Hovind Hougen, K. (1984) Ultrastructure of spirochetes isolated from Ixodes ricinus and Ixodes dammini. Yale Journal of Biology and Medicine 57, 543–548. Howe, T.R., LaQuier, F.W. and Barbour, A.G. (1986) Organization of genes encoding two outer membrane proteins of the Lyme disease agent Borrelia burgdorferi within a single transcriptional unit. Infection and Immunity 54, 207–212. Hu, C.M., Humair, P.F., Wallich, R. and Gern, L. (1997) Apodemus sp. rodents, reservoir hosts for Borrelia afzelii in an endemic area in Switzerland. Zentralblatt für Bakteriologie 285, 558–564. Hubálek, Z. and Halouzka, J. (1997) Distribution of Borrelia burgdorferi sensu lato genomic groups in Europe, a review. European Journal of Epidemiology 13, 951–957. Hudson, B.J., Stewart, M., Lennox, V.A., Fukunaga, M., Yabuki, M., Macorison, H. and Kitchener-Smith, J. (1998) Culture-positive Lyme borreliosis. Medical Journal of Australia 168, 500–502. Hyde, F.W. and Johnson, R.C. (1984) Genetic relationship of Lyme disease spirochetes to Borrelia, Treponema, and Leptospira spp. Journal of Clinical Microbiology 20, 151–154. Indest, K.J. and Philipp, M.T. (2000) DNA-binding proteins possibly involved in regulation of the post-logarithmicphase expression of lipoprotein P35 in

Molecular and Cellular Biology of B. burgdorferi

Borrelia burgdorferi. Journal of Bacteriology 182, 522–525. Indest, K.J., Ramamoorthy, R., Sole, M., Gilmore, R.D., Johnson, B.J. and Philipp, M.T. (1997) Cell-densitydependent expression of Borrelia burgdorferi lipoproteins in vitro. Infection and Immunity 65, 1165–1171. Jauris-Heipke, S., Liegl, G., Preac-Mursic, V., Rossler, D., Schwab, E., Soutschek, E., Will, G. and Wilske, B. (1995) Molecular analysis of genes encoding outer surface protein C (OspC) of Borrelia burgdorferi sensu lato: relationship to ospA genotype and evidence of lateral gene exchange of ospC. Journal of Clinical Microbiology 33, 1860–1866. Jauris-Heipke, S., Rossle, B., Wanner, G., Habermann, C., Rossler, D., Fingerle, V., Lehnert, G., Lobentanzer, R., Pradel, I., Hillenbrand, B., SchulteSpechtel, U. and Wilske, B. (1999) Osp17, a novel immunodominant outer surface protein of Borrelia afzelii: recombinant expression in Escherichia coli and its use as a diagnostic antigen for serodiagnosis of Lyme borreliosis. Medical Microbiology and Immunology 187, 213–219. Johnson, R.C., Schmid, G.P., Hyde, F.W., Steigerwalt, A.G. and Brenner, D.J. (1984) Borrelia burgdorferi sp. nov.: etiologic agent of Lyme disease. International Journal of Systematic Bacteriology 34, 496–497. Jones, J.D., Bourell, K.W., Norgard, M.V. and Radolf, J.D. (1995) Membrane topology of Borrelia burgdorferi and Treponema pallidum lipoproteins. Infection and Immunity 63, 2424–2434. Jonsson, M. and Bergström, S. (1995) Transcriptional and translational regulation of the expression of the major outer surface proteins in Lyme disease Borrelia strains. Microbiology 141, 1321–1329. Jonsson, M., Noppa, L., Barbour, A.G. and Bergström, S. (1992) Heterogeneity of outer membrane proteins in Borrelia

81

burgdorferi: comparison of osp operons of three isolates of different geographic origins. Infection and Immunity 60, 1845–1853. Jwang, B., Dewing, P., Fikrig, E. and Flavell, R.A. (1995) The hook protein of Borrelia burgdorferi, encoded by the flgE gene, is serologically recognized in Lyme disease. Clinical and Diagnostic Laboratory Immunology 2, 609–615. Katona, L.I., Beck, G. and Habicht, G.S. (1992) Purification and immunological characterization of a major low-molecular-weight lipoprotein from Borrelia burgdorferi. Infection and Immunity 60, 4995–5003. Kawabata, H., Masuzawa, T. and Yanagihara, Y. (1993) Genomic analysis of Borrelia japonica sp. nov. isolated from Ixodes ovatus in Japan. Microbiology and Immunology 37, 843–848. Kimsey, R.B. and Spielman, A. (1990) Motility of Lyme disease spirochetes in fluids as viscous as the extracellular matrix. Journal of Infectious Diseases 162, 1205–1208. Kinashi, H., Shimaji, M. and Sakai, A. (1987) Giant linear plasmids in Streptomyces which code for antibiotic biosynthesis genes. Nature 328, 454–456. Kitten, T. and Barbour, A.G. (1990) Juxtaposition of expressed variable antigen genes with a conserved telomere in the bacterium Borrelia hermsii. Proceedings of the National Academy of Sciences, USA 87, 6077–6081. Knight, S.W., Kimmel, B.J., Eggers, C.H. and Samuels, D.S. (2000) Disruption of the Borrelia burgdorferi gac gene, encoding the naturally synthesized GyrA C-terminal domain. Journal of Bacteriology 182(7), 2048–2051. Koopman, M.B., Baats, E., van Vorstenbosch, C.J., van der Zeijst, B.A. and Kusters, J.G. (1992) The periplasmic flagella of Serpulina (Treponema) hyodysenteriae are composed of two sheath proteins and three core proteins. Journal of General Microbiology 138, 2697–2706.

82

S. Bergström et al.

Kornacki, J.A. and Oliver, D.B. (1998) Lyme disease-causing Borrelia species encode multiple lipoproteins homologous to peptide-binding proteins of ABC-type transporters. Infection and Immunity 66, 4115–4122. Kryuchechnikov, V.N., Korenberg, E.I., Scherbakov, S.V., Kovalevsky Iu, V. and Levin, M.L. (1988) Identification of Borrelia isolated in the USSR from Ixodes persulcatus (Schulze) ticks. Journal of Microbiology, Epidemiology and Immunobiology 12, 41–44. Kurtenbach, K., Peacey, M., Rijpkema, S.G., Hoodless, A.N., Nuttall, P.A. and Randolph, S.E. (1998) Differential transmission of the genospecies of Borrelia burgdorferi sensu lato by game birds and small rodents in England. Applied Environmental Microbiology 64, 1169–1174. Lahdenne, P., Porcella, S.F., Hagman, K.E., Akins, D.R., Popova, T.G., Cox, D.L., Katona, L.I., Radolf, J.D. and Norgard, M.V. (1997) Molecular characterization of a 6.6-kilodalton Borrelia burgdorferi outer membrane-associated lipoprotein (lp6.6) which appears to be downregulated during mammalian infection. Infection and Immunity 65, 412–421. Lam, T.T., Nguyen, T.P., Fikrig, E. and Flavell, R.A. (1994a) A chromosomal Borrelia burgdorferi gene encodes a 22kilodalton lipoprotein, P22, that is serologically recognized in Lyme disease. Journal of Clinical Microbiology 32, 876–883. Lam, T.T., Nguyen, T.P., Montgomery, R.R., Kantor, F.S., Fikrig, E. and Flavell, R.A. (1994b) Outer surface proteins E and F of Borrelia burgdorferi, the agent of Lyme disease. Infection and Immunity 62, 290–298. Le Fleche, A., Postic, D., Girardet, K., Peter, O. and Baranton, G. (1997) Characterization of Borrelia lusitaniae sp. nov. by 16S ribosomal DNA sequence analysis. International Journal of Systematic Bacteriology 47, 921–925.

Levy, S.A., Lissman, B.A. and Ficke, C.M. (1993) Performance of a Borrelia burgdorferi bacterin in borreliosis-endemic areas. Journal of the American Veterinary Medical Association 202, 1834–1838. Lex, A., Wiesmuller, K.H., Jung, G. and Bessler, W.G. (1986) A synthetic analogue of Escherichia coli lipoprotein, tripalmitoyl pentapeptide, constitutes a potent immune adjuvant. Journal of Immunology 137, 2676–2681. Li, H., Dunn, J.J., Luft, B.J. and Lawson, C.L. (1997) Crystal structure of Lyme disease antigen outer surface protein A complexed with an Fab. Proceedings of the National Academy of Sciences, USA 94, 3584–3589. Li, M., Masuzawa, T., Takada, N., Ishiguro, F., Fujita, H., Iwaki, A., Wang, H., Wang, J., Kawabata, M. and Yanagihara, Y. (1998) Lyme disease Borrelia species in northeastern China resemble those isolated from far eastern Russia and Japan. Applied Environmental Microbiology 64, 2705–2709. Li, Z., Dumas, F., Dubreuil, D. and Jacques, M. (1993) A species-specific periplasmic flagellar protein of Serpulina (Treponema) hyodysenteriae. Journal of Bacteriology 175, 8000–8007. Liang, F.T. and Philipp, M.T. (1999) Analysis of antibody response to invariable regions of VlsE, the variable surface antigen of Borrelia burgdorferi. Infection and Immunity 67, 6702–6706. Lin, B., Short, S.A., Eskildsen, M., Klempner, M.S. and Hu, L.T. (2001) Functional testing of putative oligopeptide permease (Opp) proteins of Borrelia burgdorferi: a complementation model in opp(–) Escherichia coli. Biochimica et Biophysica Acta 1499, 222–231. Lin, Y.S., Kieser, H.M., Hopwood, D.A. and Chen, C.W. (1993) The chromosomal DNA of Streptomyces lividans 66 is linear. Molecular Microbiology 10, 923–933. Livey, I., Gibbs, C.P., Schuster, R. and Dorner, F. (1995) Evidence for lateral transfer and recombination in OspC

Molecular and Cellular Biology of B. burgdorferi

variation in Lyme disease Borrelia. Molecular Microbiology 18, 257–269. Luft, B.J., Jiang, W., Munoz, P., Dattwyler, R.J. and Gorevic, P.D. (1989) Biochemical and immunological characterization of the surface proteins of Borrelia burgdorferi. Infection and Immunity 57, 3637–3645. Magnarelli, L.A., Anderson, J.F. and Johnson, R.C. (1987) Cross-reactivity in serological tests for Lyme disease and other spirochetal infections. Journal of Infectious Diseases 156, 183–188. Marconi, R.T. and Garon, C.F. (1992a) Phylogenetic analysis of the genus Borrelia: a comparison of North American and European isolates of Borrelia burgdorferi. Journal of Bacteriology 174, 241–244. Marconi, R.T. and Garon, C.F. (1992b) Identification of a third genomic group of Borrelia burgdorferi through signature nucleotide analysis and 16S rRNA sequence determination. Journal of General Microbiology 138, 533–536. Marconi, R.T., Samuels, D.S., Landry, R.K. and Garon, C.F. (1994) Analysis of the distribution and molecular heterogeneity of the ospD gene among the Lyme disease spirochetes: evidence for lateral gene exchange. Journal of Bacteriology 176, 4572–4582. Marconi, R.T., Liveris, D. and Schwartz, I. (1995) Identification of novel insertion elements, restriction fragment length polymorphism patterns, and discontinuous 23S rRNA in Lyme disease spirochetes: phylogenetic analyses of rRNA genes and their intergenic spacers in Borrelia japonica sp. nov. and genomic group 21038 (Borrelia andersonii sp. nov.) isolates. Journal of Clinical Microbiology 33, 2427–2434. Margolis, N. and Rosa, P.A. (1993) Regulation of expression of major outer surface proteins in Borrelia burgdorferi. Infection and Immunity 61, 2207–2210. Marti Ras, N., Postic, D., Foretz, M. and

83

Baranton, G. (1997) Borrelia burgdorferi sensu stricto, a bacterial species ‘made in the U.S.A.’? International Journal of Systematic Bacteriology 47, 1112–1117. Masuzawa, T., Komikado, T., Iwaki, A., Suzuki, H., Kaneda, K. and Yanagihara, Y. (1996) Characterization of Borrelia sp. isolated from Ixodes tanuki, I. turdus, and I. columnae in Japan by restriction fragment length polymorphism of rrf (5S)–rrl (23S) intergenic spacer amplicons. FEMS Microbiology Letters 142, 77–83. Masuzawa, T., Takada, N., Kudeken, M., Fukui, T., Yano, Y., Ishiguro, F., Kawamura, Y., Imai, Y. and Ezaki, T. (2001) Borrelia sinica sp. nov., a Lyme disease-related Borrelia species isolated in China. International Journal of Systematic and Evolutionary Microbiology 51, 1817–1824. Meinhardt, F., Schaffrath, R. and Larsen, M. (1997) Microbial linear plasmids. Applied Microbiology and Biotechnology 47, 329–336. de Michelis, S., Sewell, H.S., CollaresPereira, M., Santos-Reis, M., Schouls, L.M., Benes, V., Holmes, E.C. and Kurtenbach, K. (2000) Genetic diversity of Borrelia burgdorferi sensu lato in ticks from mainland Portugal. Journal of Clinical Microbiology 38, 2128–2133. van Mierlo, P., Jacob, W. and Dockx, P. (1993) Erythema chronicum migrans: an electron-microscopic study. Dermatology 186, 306–310. Miller, J.C., Bono, J.L., Babb, K., El-Hage, N., Casjens, S. and Stevenson, B. (2000) A second allele of eppA in Borrelia burgdorferi strain B31 is located on the previously undetected circular plasmid cp9-2. Journal of Bacteriology 182, 6254–6258. Motaleb, M.A., Corum, L., Bono, J.L., Elias, A.F., Rosa, P., Samuels, D.S. and Charon, N.W. (2000) Borrelia burgdorferi periplasmic flagella have both skeletal and motility functions. Proceedings of the National Academy of Sciences, USA 97, 10899–10904.

84

S. Bergström et al.

Neubert, U., Schaller, M., Januschke, E., Stolz, W. and Schmieger, H. (1993) Bacteriophages induced by ciprofloxacin in a Borrelia burgdorferi skin isolate. Zentralblatt für Bakteriologie 279, 307–315. Noppa, L., Östberg, Y., Lavrinovicha, M. and Bergström, S. (2001) P13, an integral membrane protein of Borrelia burgdorferi, is C-terminally processed and contains surface-exposed domains. Infection and Immunity 69, 3323–3334. Norris, S.J. (1993) Polypeptides of Treponema pallidum: progress toward understanding their structural, functional, and immunologic roles. Treponema pallidum Polypeptide Research Group [published erratum appears in Microbiological Reviews (1994) 58, 291]. Microbiological Reviews 57, 750–779. Norris, S.J., Charon, N.W., Cook, R.G., Fuentes, M.D. and Limberger, R.J. (1988) Antigenic relatedness and N-terminal sequence homology define two classes of periplasmic flagellar proteins of Treponema pallidum subsp. pallidum and Treponema phagedenis. Journal of Bacteriology 170, 4072–4082. Norris, S.J., Carter, C.J., Howell, J.K. and Barbour, A.G. (1992) Low-passageassociated proteins of Borrelia burgdorferi B31: characterization and molecular cloning of OspD, a surface-exposed, plasmid-encoded lipoprotein. Infection and Immunity 60, 4662–4672. Norris, S.J., Howell, J.K., Garza, S.A., Ferdows, M.S. and Barbour, A.G. (1995) High- and low-infectivity phenotypes of clonal populations of in vitro-cultured Borrelia burgdorferi. Infection and Immunity 63, 2206–2212. Nosek, J., Tomaska, L., Fukuhara, H., Suyama, Y. and Kovac, L. (1998) Linear mitochondrial genomes: 30 years down the line. Trends in Genetics 14, 184–188. Ohnishi, J., Piesman, J. and de Silva, A.M. (2001) Antigenic and genetic heterogeneity of Borrelia burgdorferi populations

transmitted by ticks. Proceedings of the National Academy of Sciences, USA 98, 670–675. Olsen, B., Jaenson, T.G., Noppa, L., Bunikis, J. and Bergström, S. (1993) A Lyme borreliosis cycle in seabirds and Ixodes uriae ticks. Nature 362, 340–342. Olsen, B., Duffy, D.C., Jaenson, T.G., Gylfe, Å., Bonnedahl, J. and Bergström, S. (1995) Transhemispheric exchange of Lyme disease spirochetes by seabirds. Journal of Clinical Microbiology 33, 3270–3274. Olsen, G.J., Woese, C.R. and Overbeek, R. (1994) The winds of (evolutionary) change: breathing new life into microbiology. Journal of Bacteriology 176, 1–6. Pal, U., de Silva, A.M., Montgomery, R.R., Fish, D., Anguita, J., Anderson, J.F., Lobet, Y. and Fikrig, E. (2000) Attachment of Borrelia burgdorferi within Ixodes scapularis mediated by outer surface protein A. Journal of Clinical Investigation 106, 561–569. Payne, J.W. and Gilvarg, C. (1968) Size restriction on peptide utilization in Escherichia coli. Journal of Biological Chemistry 243, 6291–6299. Persing, D.H., Rutledge, B.J., Rys, P.N., Podzorski, D.S., Mitchell, P.D., Reed, K.D., Liu, B., Fikrig, E. and Malawista, S.E. (1994) Target imbalance: disparity of Borrelia burgdorferi genetic material in synovial fluid from Lyme arthritis patients. Journal of Infectious Diseases 169, 668–672. Peters, D.J. and Benach, J.L. (1997) Borrelia burgdorferi adherence and injury to undifferentiated and differentiated neural cells in vitro. Journal of Infectious Diseases 176, 470–477. Picken, R.N., Cheng, Y., Strle, F. and Picken, M.M. (1996) Patient isolates of Borrelia burgdorferi sensu lato with genotypic and phenotypic similarities of strain 25015. Journal of Infectious Diseases 174, 1112–1115. Plasterk, R.H., Simon, M.I. and Barbour, A.G. (1985) Transposition of structural

Molecular and Cellular Biology of B. burgdorferi

genes to an expression sequence on a linear plasmid causes antigenic variation in the bacterium Borrelia hermsii. Nature 318, 257–263. Porcella, S.F., Popova, T.G., Akins, D.R., Li, M., Radolf, J.D. and Norgard, M.V. (1996) Borrelia burgdorferi supercoiled plasmids encode multicopy tandem open reading frames and a lipoprotein gene family. Journal of Bacteriology 178, 3293–3307. Porcella, S.F., Fitzpatrick, C.A. and Bono, J.L. (2000) Expression and immunological analysis of the plasmid-borne mlp genes of Borrelia burgdorferi strain B31. Infection and Immunity 68, 4992–5001. Postic, D., Assous, M.V., Grimont, P.A. and Baranton, G. (1994) Diversity of Borrelia burgdorferi sensu lato evidenced by restriction fragment length polymorphism of rrf (5S)–rrl (23S) intergenic spacer amplicons. International Journal of Systematic Bacteriology 44, 743–752. Postic, D., Ras, N.M., Lane, R.S., Hendson, M. and Baranton, G. (1998) Expanded diversity among Californian Borrelia isolates and description of Borrelia bissettii sp. nov. (formerly Borrelia group DN127). Journal of Clinical Microbiology 36, 3497–3504. Probert, W.S. and Johnson, B.J. (1998) Identification of a 47 kDa fibronectinbinding protein expressed by Borrelia burgdorferi isolate B31. Molecular Microbiology 30, 1003–1015. Probert, W.S., Kim, J.H., Hook, M. and Johnson, B.J. (2001) Mapping the ligand-binding region of Borrelia burgdorferi fibronectin-binding protein BBK32. Infection and Immunity 69, 4129–4133. Pugsley, A.P. (1993) The complete general secretory pathway in Gram-negative bacteria. Microbiological Reviews 57, 50–108. Purser, J.E. and Norris, S.J. (2000) Correlation between plasmid content and infectivity in Borrelia burgdorferi. Proceedings of the National Academy of Sciences, USA 97, 13865–13870. Radolf, J.D. (1994) Role of outer membrane

85

architecture in immune evasion by Treponema pallidum and Borrelia burgdorferi. Trends in Microbiology 2, 307–311. Radolf, J.D., Bourell, K.W., Akins, D.R., Brusca, J.S. and Norgard, M.V. (1994) Analysis of Borrelia burgdorferi membrane architecture by freeze–fracture electron microscopy. Journal of Bacteriology 176, 21–31. Radolf, J.D., Goldberg, M.S., Bourell, K., Baker, S.I., Jones, J.D. and Norgard, M.V. (1995a) Characterization of outer membranes isolated from Borrelia burgdorferi, the Lyme disease spirochete. Infection and Immunity 63, 2154–2163. Radolf, J.D., Arndt, L.L., Akins, D.R., Curetty, L.L., Levi, M.E., Shen, Y., Davis, L.S. and Norgard, M.V. (1995b) Treponema pallidum and Borrelia burgdorferi lipoproteins and synthetic lipopeptides activate monocytes/macrophages. Journal of Immunology 154, 2866–2877. Restrepo, B.I. and Barbour, A.G. (1994) Antigen diversity in the bacterium B. hermsii through ‘somatic’ mutations in rearranged vmp genes. Cell 78, 867–876. Restrepo, B.I., Carter, C.J. and Barbour, A.G. (1994) Activation of a vmp pseudogene in Borrelia hermsii: an alternate mechanism of antigenic variation during relapsing fever. Molecular Microbiology 13, 287–299. Rosa, P.A., Schwan, T. and Hogan, D. (1992) Recombination between genes encoding major outer surface proteins A and B of Borrelia burgdorferi. Molecular Microbiology 6, 3031–3040. Rosey, E.L., Kennedy, M.J., Petrella, D.K., Ulrich, R.G. and Yancey, R.J. Jr (1995) Inactivation of Serpulina hyodysenteriae flaA1 and flaB1 periplasmic flagellar genes by electroporation-mediated allelic exchange. Journal of Bacteriology 177, 5959–5970. Rybchin, V.N. and Svarchevsky, A.N. (1999) The plasmid prophage N15: a linear DNA with covalently closed ends. Molecular Microbiology 33, 895–903. Sadziene, A., Thomas, D.D., Bundoc,

86

S. Bergström et al.

V.G., Holt, S.C. and Barbour, A.G. (1991) A flagella-less mutant of Borrelia burgdorferi. Structural, molecular, and in vitro functional characterization. Journal of Clinical Investigation 88, 82–92. Sadziene, A., Wilske, B., Ferdows, M.S. and Barbour, A.G. (1993a) The cryptic ospC gene of Borrelia burgdorferi B31 is located on a circular plasmid. Infection and Immunity 61, 2192–2195. Sadziene, A., Barbour, A.G., Rosa, P.A. and Thomas, D.D. (1993b) An OspB mutant of Borrelia burgdorferi has reduced invasiveness in vitro and reduced infectivity in vivo. Infection and Immunity 61, 3590–3596. Sadziene, A., Thomas, D.D. and Barbour, A.G. (1995) Borrelia burgdorferi mutant lacking Osp: biological and immunological characterization. Infection and Immunity 63, 1573–1580. Sadziene, A., Thompson, P.A. and Barbour, A.G. (1996) A flagella-less mutant of Borrelia burgdorferi as a live attenuated vaccine in the murine model of Lyme disease. Journal of Infectious Diseases 173, 1184–1193. Samuels, D.S., Mach, K.E. and Garon, C.F. (1994a) Genetic transformation of the Lyme disease agent Borrelia burgdorferi with coumarin-resistant gyrB. Journal of Bacteriology 176, 6045–6049. Samuels, D.S., Marconi, R.T., Huang, W.M. and Garon, C.F. (1994b) gyrB mutations in coumermycin A1-resistant Borrelia burgdorferi. Journal of Bacteriology 176, 3072–3075. Sankaran, K., Gupta, S.D. and Wu, H.C. (1995) Modification of bacterial lipoproteins. Methods in Enzymology 250, 683–697. Sartakova, M., Dobrikova, E. and Cabello, F.C. (2000) Development of an extrachromosomal cloning vector system for use in Borrelia burgdorferi. Proceedings of the National Academy of Sciences, USA 97, 4850–4855. Schmid, G.P., Steigerwalt, A.G., Johnson, S.E., Barbour, A.G., Steere, A.C.,

Robinson, I.M. and Brenner, D.J. (1984) DNA characterization of the spirochete that causes Lyme disease. Journal of Clinical Microbiology 20, 155–158. Schoenfeld, R., Araneo, B., Ma, Y., Yang, L.M. and Weis, J.J. (1992) Demonstration of a B-lymphocyte mitogen produced by the Lyme disease pathogen, Borrelia burgdorferi. Infection and Immunity 60, 455–464. Schwan, T.G. and Burgdorfer, W. (1987) Antigenic changes of Borrelia burgdorferi as a result of in vitro cultivation. Journal of Infectious Diseases 156, 852–853. Schwan, T.G. and Piesman, J. (2000) Temporal changes in outer surface proteins A and C of the Lyme disease-associated spirochete, Borrelia burgdorferi, during the chain of infection in ticks and mice. Journal of Clinical Microbiology 38, 382–388. Schwan, T.G., Burgdorfer, W. and Garon, C.F. (1988) Changes in infectivity and plasmid profile of the Lyme disease spirochete, Borrelia burgdorferi, as a result of in vitro cultivation. Infection and Immunity 56, 1831–1836. Schwan, T.G., Piesman, J., Golde, W.T., Dolan, M.C. and Rosa, P.A. (1995) Induction of an outer surface protein on Borrelia burgdorferi during tick feeding. Proceedings of the National Academy of Sciences, USA 92, 2909–2913. Seshu, J. and Skare, J.T. (2000) The many faces of Borrelia burgdorferi. Journal of Molecular Microbiology and Biotechnology 2, 463–472. Shi, W., Yang, Z., Geng, Y., Wolinsky, L.E. and Lovett, M.A. (1998) Chemotaxis in Borrelia burgdorferi. Journal of Bacteriology 180, 231–235. Shoberg, R.J. and Thomas, D.D. (1993) Specific adherence of Borrelia burgdorferi extracellular vesicles to human endothelial cells in culture. Infection and Immunity 61, 3892–3900. Shoberg, R.J. and Thomas, D.D. (1995) Borrelia burgdorferi vesicle production

Molecular and Cellular Biology of B. burgdorferi

occurs via a mechanism independent of immunoglobulin M involvement. Infection and Immunity 63, 4857–4861. Sigal, L.H., Zahradnik, J.M., Lavin, P., Patella, S.J., Bryant, G., Haselby, R., Hilton, E., Kunkel, M., Adler-Klein, D., Doherty, T., Evans, J. and Malawista, S.E. (1998) A vaccine consisting of recombinant Borrelia burgdorferi outer-surface protein A to prevent Lyme disease. Recombinant OuterSurface Protein A Lyme Disease Vaccine Study Consortium. New England Journal of Medicine 339, 216–222. de Silva, A.M., Telford, S.R., 3rd, Brunet, L.R., Barthold, S.W. and Fikrig, E. (1996) Borrelia burgdorferi OspA is an arthropod-specific transmission-blocking Lyme disease vaccine. Journal of Experimental Medicine 183, 271–275. de Silva, A.M., Zeidner, N.S., Zhang, Y., Dolan, M.C., Piesman, J. and Fikrig, E. (1999) Influence of outer surface protein A antibody on Borrelia burgdorferi within feeding ticks. Infection and Immunity 67, 30–35. Simpson, W.J., Garon, C.F. and Schwan, T.G. (1990) Analysis of supercoiled circular plasmids in infectious and noninfectious Borrelia burgdorferi. Microbial Pathogenicity 8, 109–118. Simpson, W.J., Schrumpf, M.E., Hayes, S.F. and Schwan, T.G. (1991a) Molecular and immunological analysis of a polymorphic periplasmic protein of Borrelia burgdorferi. Journal of Clinical Microbiology 29, 1940–1948. Simpson, W.J., Burgdorfer, W., Schrumpf, M.E., Karstens, R.H. and Schwan, T.G. (1991b) Antibody to a 39-kilodalton Borrelia burgdorferi antigen (P39) as a marker for infection in experimentally and naturally inoculated animals. Journal of Clinical Microbiology 29, 236–243. Skare, J.T., Shang, E.S., Foley, D.M., Blanco, D.R., Champion, C.I., Mirzabekov, T., Sokolov, Y., Kagan, B.L., Miller, J.N. and Lovett, M.A.

87

(1995) Virulent strain associated outer membrane proteins of Borrelia burgdorferi. Journal of Clinical Investigation 96, 2380–2392. Skare, J.T., Champion, C.I., Mirzabekov, T.A., Shang, E.S., Blanco, D.R., Erdjument-Bromage, H., Tempst, P., Kagan, B.L., Miller, J.N. and Lovett, M.A. (1996) Porin activity of the native and recombinant outer membrane protein Oms28 of Borrelia burgdorferi. Journal of Bacteriology 178, 4909–4918. Skare, J.T., Mirzabekov, T.A., Shang, E.S., Blanco, D.R., ErdjumentBromage, H., Bunikis, J., Bergström, S., Tempst, P., Kagan, B.L., Miller, J.N. and Lovett, M.A. (1997) The Oms66 (p66) protein is a Borrelia burgdorferi porin. Infection and Immunity 65, 3654–3661. de Souza, M.S., Fikrig, E., Smith, A.L., Flavell, R.A. and Barthold, S.W. (1992) Nonspecific proliferative responses of murine lymphocytes to Borrelia burgdorferi antigens. Journal of Infectious Diseases 165, 471–478. Steere, A.C., Sikand, V.K., Meurice, F., Parenti, D.L., Fikrig, E., Schoen, R.T., Nowakowski, J., Schmid, C.H., Laukamp, S., Buscarino, C. and Krause, D.S. (1998) Vaccination against Lyme disease with recombinant Borrelia burgdorferi outer-surface lipoprotein A with adjuvant. New England Journal of Medicine 339, 209–215. Steere, A.C., Gross, D., Meyer, A.L. and Huber, B.T. (2001) Autoimmune mechanisms in antibiotic treatment-resistant Lyme arthritis. Journal of Autoimmunity 16, 263–268. Stevenson, B., Tilly, K. and Rosa, P.A. (1996) A family of genes located on four separate 32-kilobase circular plasmids in Borrelia burgdorferi B31. Journal of Bacteriology 178, 3508–3516. Stevenson, B., Bono, J.L., Schwan, T.G. and Rosa, P. (1998) Borrelia burgdorferi Erp proteins are immunogenic in mammals infected by tick bite, and their

88

S. Bergström et al.

synthesis is inducible in cultured bacteria. Infection and Immunity 66, 2648–2654. Stevenson, B., Zuckert, W.R. and Akins, D.R. (2000a) Repetition, conservation, and variation: the multiple cp32 plasmids of Borrelia species. Journal of Molecular Microbiology and Biotechnology 2, 411–422. Stevenson, B., Porcella, S.F., Oie, K.L., Fitzpatrick, C.A., Raffel, S.J., Lubke, L., Schrumpf, M.E. and Schwan, T.G. (2000b) The relapsing fever spirochete Borrelia hermsii contains multiple, antigen-encoding circular plasmids that are homologous to the cp32 plasmids of Lyme disease spirochetes. Infection and Immunity 68, 3900–3908. Stewart, P.E., Thalken, R., Bono, J.L. and Rosa, P. (2001) Isolation of a circular plasmid region sufficient for autonomous replication and transformation of infectious Borrelia burgdorferi. Molecular Microbiology 39, 714–721. Strle, F., Picken, R.N., Cheng, Y., Cimperman, J., Maraspin, V., LotricFurlan, S., Ruzic-Sabljic, E. and Picken, M.M. (1997) Clinical findings for patients with Lyme borreliosis caused by Borrelia burgdorferi sensu lato with genotypic and phenotypic similarities to strain 25015. Clinical Infectious Disease 25, 273–280. Suk, K., Das, S., Sun, W., Jwang, B., Barthold, S.W., Flavell, R.A. and Fikrig, E. (1995) Borrelia burgdorferi genes selectively expressed in the infected host. Proceedings of the National Academy of Sciences, USA 92, 4269–4273. Sung, S.Y., McDowell, J.V., Carlyon, J.A. and Marconi, R.T. (2000) Mutation and recombination in the upstream homology box-flanked ospE-related genes of the Lyme disease spirochetes result in the development of new antigenic variants during infection. Infection and Immunity 68, 1319–1327. Swancutt, M.A., Riley, B.S., Radolf, J.D. and Norgard, M.V. (1989) Molecular

characterization of the pathogen-specific, 34-kilodalton membrane immunogen of Treponema pallidum. Infection and Immunity 57, 3314–3323. Takayama, K., Rothenberg, R.J. and Barbour, A.G. (1987) Absence of lipopolysaccharide in the Lyme disease spirochete, Borrelia burgdorferi. Infection and Immunity 55, 2311–2313. Theisen, M., Borre, M., Mathiesen, M.J., Mikkelsen, B., Lebech, A.M. and Hansen, K. (1995) Evolution of the Borrelia burgdorferi outer surface protein OspC. Journal of Bacteriology 177, 3036–3044. Thomas, D.D. and Comstock, L.E. (1989) Interaction of Lyme disease spirochetes with cultured eucaryotic cells. Infection and Immunity 57, 1324–1336. Tilly, K., Casjens, S., Stevenson, B., Bono, J.L., Samuels, D.S., Hogan, D. and Rosa, P. (1997) The Borrelia burgdorferi circular plasmid cp26: conservation of plasmid structure and targeted inactivation of the ospC gene. Molecular Microbiology 25(2), 361–373. Tilly, K., Lubke, L. and Rosa, P. (1998) Characterization of circular plasmid dimers in Borrelia burgdorferi. Journal of Bacteriology 180(21), 5676–5681. Tilly, K., Elias, A.F., Bono, J.L., Stewart, P. and Rosa, P. (2000) DNA exchange and insertional inactivation in spirochetes. Journal of Molecular Microbiology and Biotechnology 2, 433–442. Tilly, K., Elias, A.F., Errett, J. Fischer, E. Iyer, R., Schwartz, I., Bono, J.L. and Rosa, P. (2001) Genetics and regulation of chitobiose utilization in Borrelia burgdorferi. Journal of Bacteriology 183(19), 5544–5553. Tomb, J.F., White, O., Kerlavage, A.R., Clayton, R.A., Sutton, G.G., Fleischmann, R.D., Ketchum, K.A., Klenk, H.P., Gill, S., Dougherty, B.A., Nelson, K., Quackenbush, J., Zhou, L., Kirkness, E.F., Peterson, S., Loftus, B., Richardson, D., Dodson, R., Khalak, H.G., Glodek, A., McKenney, K.,

Molecular and Cellular Biology of B. burgdorferi

Fitzegerald, L.M., Lee, N., Adams, M.D., Hickey, E.K., Berg, D.E., Gocayne, J.D., Utterback, T.R., Peterson, J.D., Kelley, J.M., Cotton, M.D., Weidman, J.M., Fujii, C., Bowman, C., Watthey, L., Wallin, E., Hayes, W.S., Borodovsky, M., Karp, P.D., Smith, H.O., Fraser, C.M. and Venter, J.C. (1997) The complete genome sequence of the gastric pathogen Helicobacter pylori. Nature 388, 539–547. Trollmo, C., Meyer, A.L., Steere, A.C., Hafler, D.A. and Huber, B.T. (2001) Molecular mimicry in Lyme arthritis demonstrated at the single cell level: LFA-1 alpha L is a partial agonist for outer surface protein A-reactive T cells. Journal of Immunology 166, 5286–5291. Trueba, G.A., Bolin, C.A. and Zuerner, R.L. (1992) Characterization of the periplasmic flagellum proteins of Leptospira interrogans. Journal of Bacteriology 174, 4761–4768. Trueba, G.A., Old, I.G., Saint Girons, I. and Johnson, R.C. (1997) A cheA and cheW operon in Borrelia burgdorferi. Research in Microbiology 148, 191–200. Valsangiacomo, C., Balmelli, T. and Piffaretti, J.C. (1997) A phylogenetic analysis of Borrelia burgdorferi sensu lato based on sequence information from the hbb gene, coding for a histone-like protein. International Journal of Systematic Bacteriology 47, 1–10. Volff, J.N. and Altenbuchner, J. (2000) A new beginning with new ends: linearisation of circular chromosomes during bacterial evolution. FEMS Microbiology Letters 186, 143–150. Walker, E.M., Borenstein, L.A., Blanco, D.R., Miller, J.N. and Lovett, M.A. (1991) Analysis of outer membrane ultrastructure of pathogenic Treponema and Borrelia species by freeze–fracture electron microscopy. Journal of Bacteriology 173, 5585–5588. Wallich, R., Moter, S.E., Simon, M.M.,

89

Ebnet, K., Heiberger, A. and Kramer, M.D. (1990) The Borrelia burgdorferi flagellum-associated 41-kilodalton antigen (flagellin): molecular cloning, expression, and amplification of the gene. Infection and Immunity 58, 1711–1719. Wallich, R., Simon, M.M., Hofmann, H., Moter, S.E., Schaible, U.E. and Kramer, M.D. (1993) Molecular and immunological characterization of a novel polymorphic lipoprotein of Borrelia burgdorferi. Infection and Immunity 61, 4158–4166. Wang, G., van Dam, A.P., Le Fleche, A., Postic, D., Peter, O., Baranton, G., de Boer, R., Spanjaard, L. and Dankert, J. (1997) Genetic and phenotypic analysis of Borrelia valaisiana sp. nov. (Borrelia genomic groups VS116 and M19). International Journal of Systematic Bacteriology 47, 926–932. Wang, G., van Dam, A.P. and Dankert, J. (1999a) Phenotypic and genetic characterization of a novel Borrelia burgdorferi sensu lato isolate from a patient with Lyme borreliosis. Journal of Clinical Microbiology 37, 3025–3028. Wang, G., van Dam, A.P. and Dankert, J. (1999b) Evidence for frequent OspC gene transfer between Borrelia valaisiana sp. nov. and other Lyme disease spirochetes. FEMS Microbiology Letters 177, 289–296. Wang, G., van Dam, A.P., Schwartz, I. and Dankert, J. (1999c) Molecular typing of Borrelia burgdorferi sensu lato: taxonomic, epidemiological, and clinical implications. Clinical Microbiology Reviews 12, 633–653. Wang, G., van Dam, A.P. and Dankert, J. (2000) Two distinct ospA genes among Borrelia valaisiana strains. Research in Microbiology 151, 325–331. Wang, I.N., Dykhuizen, D.E., Qiu, W., Dunn, J.J., Bosler, E.M. and Luft, B.J. (1999) Genetic diversity of ospC in a local population of Borrelia burgdorferi sensu stricto. Genetics 151, 15–30.

90

S. Bergström et al.

Wayne, L.G., Brenner, D.J., Colwell, R.R., Grimont, P.A.D., Kandler, O., Krichevsky, M.I., Moore, L.H., Moore, W.E.C., Murray, R.G.E., Stackebrandt, E., Starr, M.P. and Truper, H.G. (1987) Report of the Ad Hoc Committee on reconciliation of approaches to bacterial systematics. International Journal of Systematic Bacteriology 37, 463–464. Wheeler, C.M., Garcia Monco, J.C., Benach, J.L., Golightly, M.G., Habicht, G.S. and Steere, A.C. (1993) Nonprotein antigens of Borrelia burgdorferi. Journal of Infectious Diseases 167, 665–674. Wilske, B., Preac-Mursic, V., Gobel, U.B., Graf, B., Jauris, S., Soutschek, E., Schwab, E. and Zumstein, G. (1993) An OspA serotyping system for Borrelia burgdorferi based on reactivity with monoclonal antibodies and OspA sequence analysis. Journal of Clinical Microbiology 31, 340–350. Wilske, B., Jauris-Heipke, S., Lobentanzer, R., Pradel, I., Preac-Mursic, V., Rössler, D., Soutschek, E. and Johnson, R.C. (1995) Phenotypic analysis of outer surface protein C (OspC) of Borrelia burgdorferi sensu lato by monoclonal antibodies: relationship to genospecies and OspA serotype. Journal of Clinical Microbiology 33, 103–109. Wilske, B., Busch, U., Fingerle, V., JaurisHeipke, S., Preac Mursic, V., Rössler, D. and Will, G. (1996) Immunological and molecular variability of OspA and OspC. Implications for Borrelia vaccine development. Infection 24, 208–212. Wlodarzyk, M. and Nowicka, B. (1988) Preliminary evidence for the linear nature of Thiobacillus versutus pTAV2 plasmid. FEMS Microbiology Letters 55, 125–128. Woese, C.R. (1987) Bacterial evolution. Microbiological Reviews 51, 221–271. Woese, C.R., Olsen, G.J., Ibba, M. and Soll, D. (2000) Aminoacyl-tRNA synthetases, the genetic code, and the evolutionary process. Microbiology and

Molecular Biology Reviews 64, 202–236. Xu, Y., Kodner, C., Coleman, L. and Johnson, R.C. (1996) Correlation of plasmids with infectivity of Borrelia burgdorferi sensu stricto type strain B31. Infection and Immunity 64, 3870–3876. Yang, X., Popova, T.G., Hagman, K.E., Wikel, S.K., Schoeler, G.B., Caimano, M.J., Radolf, J.D. and Norgard, M.V. (1999) Identification, characterization, and expression of three new members of the Borrelia burgdorferi Mlp (2.9) lipoprotein gene family. Infection and Immunity 67, 6008–6018. Yang, X., Goldberg, M.S., Popova, T.G., Schoeler, G.B., Wikel, S.K., Hagman, K.E. and Norgard, M.V. (2000) Interdependence of environmental factors influencing reciprocal patterns of gene expression in virulent Borrelia burgdorferi. Molecular Microbiology 37, 1470–1479. Zhang, J.R. and Norris, S.J. (1998a) Kinetics and in vivo induction of genetic variation of vlsE in Borrelia burgdorferi. Infection and Immunity 66, 3689–3697. Zhang, J.R. and Norris, S.J. (1998b) Genetic variation of the Borrelia burgdorferi gene vlsE involves cassette-specific, segmental gene conversion. Infection and Immunity 66, 3698–3704. Zhang, J.R., Hardham, J.M., Barbour, A.G. and Norris, S.J. (1997) Antigenic variation in Lyme disease Borreliae by promiscuous recombination of VMPlike sequence cassettes. Cell 89, 275–285. Zhang, Q.Y., DeRyckere, D., Lauer, P. and Koomey, M. (1992) Gene conversion in Neisseria gonorrhoeae: evidence for its role in pilus antigenic variation. Proceedings of the National Academy of Sciences, USA 89, 5366–5370. Zuckert, W.R. and Meyer, J. (1996) Circular and linear plasmids of Lyme disease spirochetes have extensive homology: characterization of a repeated DNA element. Journal of Bacteriology 178, 2287–2298.

Vectors of Borrelia burgdorferi sensu lato

4

Lars Eisen and Robert S. Lane Department of Environmental Science, Policy and Management, University of California, Berkeley, CA 94720, USA

Introduction Lyme borreliosis (LB) is the most common arthropod-borne human disease in temperate regions of the northern hemisphere, with more than 16,800 cases occurring in the USA alone in 1998 (CDC, 2000). The primary causative agents of LB include three of the 11 genospecies currently comprising Borrelia burgdorferi sensu lato (s.l.), i.e. B. burgdorferi sensu stricto (s.s.), Borrelia afzelii and Borrelia garinii. These spirochaetal bacteria are maintained in enzootic spirochaete–tick vector–vertebrate cycles and incidental human exposure occurs when enzootic tick vectors bridge the gap from such maintenance cycles by biting humans. The primary bridging vectors of LB spirochaetes to humans include four tick species of the Ixodes ricinus complex (Acari: Ixodidae: Ixodes: subgenus Ixodes; Filippova, 1999; Keirans et al., 1999), i.e. the castor bean tick, I. ricinus, and the taiga tick, Ixodes persulcatus, in Eurasia and the western black-legged tick, Ixodes pacificus, and the black-legged tick, Ixodes scapularis, in North America (Fig. 4.1; Table 4.1). The above-mentioned tick species also include the primary vectors of the disease agents causing tick-borne encephalitis, human babesiosis and human granulocytic ehrlichiosis (see Table 4.1). The main goals of this chapter are to summarize the results of experimental studies of vector competence for B. burgdorferi s.l., give a general description of tick–B. burgdorferi s.l. relationships, provide an outline of the biology of the primary bridging vectors of LB spirochaetes and examine how differences in the ecology of these ticks affect their vector potential. Our literature review, concluded by August 2001, is restricted to the seminal reference(s) and examples of recent studies. © 2002 CAB International. Lyme Borreliosis: Biology, Epidemiology and Control (eds J. Gray, O. Kahl, R.S. Lane and G. Stanek)

91

92

L. Eisen and R.S. Lane

Table 4.1. Human disease agents transmitted by the primary bridging vectors of Lyme borreliosis spirochaetes.

Ixodes vectora

Geographical distribution

I. persulcatus I. pacificus I. scapularis I. ricinus

Far-eastern Europe, Asia Western North America Eastern North America Europe, far-western Asia, far-northern Africa

Bb ssb B afzb Xf

Pathogen TBE- Bab B garb virusc divd X

X X X

X Xg

X

X

Bab Ana micd phae

X

X X

X X X

a

I. scapularis includes records for I. dammini (Oliver et al., 1993a). Bb ss, Borrelia burgdorferi sensu stricto; B afz, B. afzelii; B gar, B. garinii; causative agents of Lyme borreliosis. c TBE-virus, tick-borne encephalitis subgroup viruses; causative agents of TBE in Europe (European or western TBE), Asia (Far-eastern TBE or Russian spring–summer encephalitis), and North America (Powassan virus). See reviews by Korenberg (1994) and Nuttall and Labuda (1994). d Bab div, Babesia divergens; Bab mic, Babesia microti; primary causative agents of human babesiosis. See reviews by Spielman et al. (1985) and Kjemtrup and Conrad (2000). e Ana pha Anaplasma (Ehrlichia) phagocytophila (Dulmer et al., 2001); causative agent of human granulocytic ehrlichiosis. See reviews by Woldehiwet (1983) and Dumler and Bakken (1998). f Vector competence indicated by presence of B. afzelii in I. persulcatus and rodents (e.g. Postic et al., 1997; Alekseev et al., 2001); experimental confirmation of vector status lacking. g I. scapularis is a confirmed vector of Powassan virus (which includes a genotype called deer tick virus; Telford et al., 1997) (Costero and Grayson, 1996), but Ixodes cookei may be the primary bridging vector to humans (Artsob, 1988). b

Vector Competence for B. burgdorferi s.l. A literature review revealed that natural infections with B. burgdorferi s.l. have been recorded from at least 25 species of Ixodes ticks, some 15 other ixodid tick species in six genera (Amblyomma, Boophilus, Dermacentor, Haemaphysalis, Hyalomma, Rhipicephalus) and two argasid tick species representing two genera (Argas, Ornithodoros). Some Ixodes ticks appear to be associated with one or a few B. burgdorferi s.l. genospecies, whereas I. ricinus harbours at least five genospecies (reviewed by Baranton et al., 1998; Wang et al., 1999). In some studies, B. burgdorferi s.l. were observed in partially or fully fed ticks removed from hosts, which demonstrates only that the ticks ingested infected body fluids. In others, B. burgdorferi s.l. were detected in questing ticks, indicating that spirochaetes ingested by larvae or nymphs survived the moult and were passed to the resulting nymphs or adults (transstadial passage). However, neither

Vectors of B. burgdorferi

93

finding demonstrates that the ticks are vectors for B. burgdorferi s.l. As used in this chapter, vector competence for B. burgdorferi s.l. means that a tick species can acquire spirochaetes when feeding on an infective host, pass them transstadially and, subsequently, transmit the spirochaetes to a susceptible host while feeding. However, a tick species may be a vector for B. burgdorferi s.l. but only make a minor contribution to enzootic spirochaete–tick vector–vertebrate cycles (referred to as enzootic cycles) or as a bridging vector to humans. Therefore, vector potential (taking into account any factors that increase or decrease vector efficiency; (see Kahl et al., Chapter 2) will be used to assess the importance of tick species/populations as enzootic vectors or bridging vectors to humans. Vector competence for B. burgdorferi s.l. has been experimentally confirmed for 12 tick species. These include six of the 14 members of the I. ricinus complex (I. affinis, I. jellisoni, I. pacificus, I. persulcatus, I. ricinus, I. scapularis) and six other Ixodes spp. (I. angustus, I. dentatus, I. hexagonus, I. minor, I. muris, I. spinipalpis) (Table 4.2). Published vector competence studies have included only four B. burgdorferi s.l. genospecies (B. burgdorferi s.s., B. afzelii, B. bissettii, B. garinii) (see Table 4.2). In contrast to most other Borrelia spp. spirochaetes, several B. burgdorferi s.l. genospecies appear to have a low vector specificity; vectors for B. burgdorferi s.s., for example, include at least nine Ixodes spp. (Sonenshine, 1993; Table 4.2). In addition to B. burgdorferi s.l., I. scapularis was recently demonstrated to be naturally infected with and a vector of a hitherto unknown relapsing fever group spirochaete closely related to Borrelia miyamotoi (Scoles et al., 2001), a spirochaete previously detected in I. persulcatus from Japan (Fukunaga et al., 1995). I. pacificus, I. persulcatus, I. ricinus and I. scapularis are the primary bridging vectors of LB spirochaetes to humans in temperate regions of the northern hemisphere (Fig. 4.1, Table 4.1). There is also evidence indicating that I. angustus, a tick species occurring in moist, cool habitats in North America and eastern Asia, acts as a secondary bridging vector (Damrow et al., 1989). The majority of the remaining confirmed vectors feed primarily on rodents and/or lagomorphs (i.e. I. dentatus, I. jellisoni, I. muris and I. spinipalpis in North America and I. minor in North and South America), whereas I. affinis infests a wide variety of mammals in North and South America, and I. hexagonus is found on various medium-sized mammals in Europe and north-western Africa (Jaenson et al., 1994; Durden and Keirans, 1996). Their host preferences render these tick species unlikely to act as bridging vectors to humans but some of them probably have a significant impact on enzootic spirochaete maintenance. In addition to the experimentally confirmed vectors, the presence of spirochaetes in ticks and their primary hosts suggest vector competence for B. burgdorferi s.l. of several other Ixodes spp. These include the cosmopolitan seabird tick I. uriae (Olsen et al., 1995), the Eurasian rodent tick I. trianguliceps (Postic et al., 1997) and the Asian ticks I. columnae, I. granulatus, I. nipponensis, I. tanuki and I. turdus (e.g. Fukunaga et al., 1996; Masuzawa et al., 1999). Seven tick species evaluated for vector competence appear unable to transmit B. burgdorferi s.l. (Amblyomma americanum, Dermacentor andersoni, D. occidentalis, D. variabilis, Ixodes cookei, I. holocyclus, I. ovatus; Table 4.3). In most cases these ticks

Ixodes tick speciesa

a

affinis angustus dentatus hexagonus jellisoni minor muris pacificus persulcatus ricinus scapularis spinipalpis

Borrelia burgdorferi sensu stricto

Borrelia garinii

SI-1 CA4 Wildd B31 CA404, 409, 445 SH-2-82 B31, SH-2-82 JD1, CA4 Wild Wild, NC-2, MI-129 Wild

Wild, ZS7 Wild, PGau B31, JD1, SI-1, SH-2-82 Pgau B31

SCW-30 CA589,592

MI-6, SCW-30 Wild, CA589, 592

JEM6 Wilde, VS286 VS286

Selected referencesc 1 2 3 4 5 1 1, 6 7–8, 24 9–10 11–14 1, 13, 15–21 21–24

I. scapularis includes records for I. dammini (Oliver et al., 1993a) and I. spinipalpis for I. neotomae (Norris et al., 1997). In some cases the genospecies was not given in the original study, but could be elucidated by later genetic studies of the isolate used. Isolates not characterized genetically, to our knowledge, are listed as B. burgdorferi sensu lato. c Studies where transmission of spirochaetes was accomplished by partially fed ticks, infected by artificial feeding or partial blood meals on infected hosts, are excluded; (1) Oliver, 1996; J.H. Oliver Jr, Georgia Southern University, 2001, personal communication; (2) Peavey et al., 2000; (3) Telford and Spielman, 1989a; (4) Gern et al., 1991; (5) Lane et al., 1999; (6) Dolan et al., 2000; (7) Lane et al., 1994; (8) Peavey and Lane, 1995; (9) Nakao and Sato, 1996; (10) Sato and Nakao, 1997; (11) Burgdorfer et al., 1983; (12) Gern and Rais, 1996; (13) Dolan et al., 1998; (14) Hu et al., 2001; (15) Burgdorfer et al., 1982; (16) Piesman and Sinsky, 1988; (17) Ryder et al., 1992; (18) Oliver et al., 1993b; (19) Sanders and Oliver, 1995; (20) Piesman and Happ, 1997; (21) Dolan et al., 1997; (22) Brown and Lane, 1992; (23) Burkot et al., 2000; (24) L. Eisen and R.S. Lane, unpublished data. d Spirochaetes originating from field-captured ticks or mammals, no isolate designation given. e Including B. garinii OspA serotype 4 (Hu et al., 2001). b

L. Eisen and R.S. Lane

I. I. I. I. I. I. I. I. I. I. I. I.

Borrelia burgdorferi sensu lato

Genospecies and isolate used b Borrelia Borrelia afzelii bissettii

94

Table 4.2. Tick species experimentally confirmed as vectors of Borrelia burgdorferi sensu lato.

I. persulcatus

I. scapularis

I. ricinus

95

Fig. 4.1. The geographical distributions of the primary vectors of Lyme borreliosis spirochaetes. (Designed by B. Kaye.)

Vectors of B. burgdorferi

I. pacificus

Genospecies and isolate useda Borrelia burgdorferi sensu stricto

Borrelia bissettii

Amblyomma americanum NC-2 Dermacentor andersonii Wild Dermacentor occidentalis CA13, 79, 85 Dermacentor variabilis NC-2, MI-119, 129 Ixodes cookei LI-231 Ixodes holocyclus Ixodes ovatus

JD1, SI-1, SH-2-82 B31 CA4, 5, 11 JD1, SI-1 SH-2-82 JD1

MI-6

a

Isolates demonstrated to be: Borrelia garinii

MI-6 JEM4, 5, 6

Acquired

Passed transstadially Transmitted

Referencesb

+/–c + + +/–e + + +

+/–c – – +/–e +/–f – –

1–7 8 9–11 1–3, 6–7 4, 12 13 14

– – –d –e – NTg NTg

Examples of isolates used. In some cases the genospecies was not given in the original study but could be elucidated from later genetic studies of the isolate used. Isolates not characterized genetically, to our knowledge, are listed as B. burgdorferi sensu lato. b (1) Piesman and Sinsky, 1988; (2) Mather and Mather, 1990; (3) Mukolwe et al., 1992; (4) Ryder et al., 1992; (5) Oliver et al., 1993b; (6) Sanders and Oliver, 1995; (7) Piesman and Happ, 1997; (8) Dolan et al., 1997; (9) Brown and Lane, 1992; (10) Lane et al., 1994; (11) Li and Lane, 1996; (12) Barker et al., 1993; (13) Piesman and Stone, 1991; (14) Nakao and Miyamoto, 1994. c Acquisition of spirochaetes during feeding was observed in three of five studies examining newly engorged larvae. Transstadial transmission was recorded by Ryder et al. (1992) (one of 60 examined nymphs infected with SH-2-82), but absent in four other studies. d Examined with artificially infected ticks; transmission of B. burgdorferi sensu stricto (CA4) from infected adult ticks to rabbits was observed for I. pacificus but not for D. occidentalis (Li and Lane, 1996). e Acquisition of spirochaetes during feeding was observed in three of four studies examining newly engorged larvae. Transstadial transmission was recorded by Piesman and Happ (1997) for isolates NC-2 (one infected nymph) and MI-119 (two infected nymphs), but absent in three other studies. Spirochaetes were not isolated from any host exposed to D. variabilis in three studies, although one of six exposed rabbits reportedly seroconverted in one of these studies (Mukolwe et al., 1992). f Isolate LI-231 was passed transstadially to 4–16% of nymphs fed as larvae on various hosts (Barker et al., 1993), whereas no transstadial passage was recorded for isolate SH-2-82 (Ryder et al., 1992). g NT, not tested.

L. Eisen and R.S. Lane

Borrelia burgdorferi sensu lato

Tick species

96

Table 4.3. Tick species apparently unable to serve as vectors of Borrelia burgdorferi sensu lato.

Vectors of B. burgdorferi

97

acquired B. burgdorferi s.l. when feeding on infected hosts but transstadial passage was rare or absent, and there was no evidence of spirochaete transmission during feeding. Haemaphysalis leporispalustris also appears to be a poor potential vector of B. burgdorferi s.l.; spirochaetes were absent from 24 unfed nymphs fed as larvae on wild-caught rabbits, whereas approximately 80% of unfed I. dentatus nymphs fed as larvae on the same individual hosts were infected (Telford and Spielman, 1989b). Until vector competence has been demonstrated for Amblyomma, Argas, Boophilus, Dermacentor, Haemaphysalis, Hyalomma, Ornithodoros or Rhipicephalus tick species, they should not be assumed to serve as bridging vectors for B. burgdorferi s.l. to humans, or to play any role in the enzootic maintenance of these spirochaetes. Also, the apparent vector incompetence of I. cookei for B. burgdorferi s.s. and I. ovatus for B. garinii (Table 4.3) underscores the need for experimental verification of Ixodes species suspected to be vectors for B. burgdorferi s.l. In addition to ticks, B. burgdorferi s.l. have been detected in numerous insect species, including mosquitoes, tabanid flies and fleas (e.g. Magnarelli and Anderson, 1988; Hubálek et al., 1998). There are some published accounts of LB following insect bites but no evidence that insects can serve as vectors for LB spirochaetes has been presented. Although insects may prove incompetent to serve as vectors for B. burgdorferi s.l., the possibility of occasional mechanical transmission should not be discounted. For example, people could be exposed to B. burgdorferi s.l. by crushing an infected insect and thereby introducing spirochaetes into the bite site or through broken skin. Problems in the design of experimental vector-competence studies are discussed by Kahl et al. (Chapter 2) and were previously addressed by Lane (1994) and Randolph and Nuttall (1994). Main problem areas include usage of cultureadapted spirochaete strains, transmission of the disease agent by tick-bite versus needle inoculation, choice of tick host and disparity in the geographical origin of the ticks and the spirochaetal isolates used.

Tick–Borrelia Relationships Recent studies have provided new insights into the dynamics at the tick–B. burgdorferi s.l. interface, as well as possible explanations for vector incompetence. Spirochaetes ingested during a blood meal multiply rapidly within the midgut of I. scapularis larvae during the first 2 weeks post-repletion but their numbers drop dramatically during the subsequent pre-moult period (Piesman et al., 1990). In moulted unfed Ixodes spp. ticks the spirochaetes are generally restricted to the midgut; during a blood meal the spirochaetes multiply in the gut before they escape through the gut wall into the haemocoel and migrate to and invade the salivary glands (e.g. Ribeiro et al., 1987; Gern et al., 1990; de Silva and Fikrig, 1995). B. burgdorferi s.s. and B. afzelii downregulate expression of outer surface protein A (OspA) and upregulate OspC during tick feeding (e.g. Schwan et al., 1995; de Silva et al., 1996; Leuba-Garcia et al., 1998). This change in surface protein profile may constitute the mechanism enabling the spirochaetes to escape

98

L. Eisen and R.S. Lane

the tick midgut and/or invade the salivary glands; OspA mediates binding of spirochaete–spirochaete and spirochaete–I. scapularis tick gut (Pal et al., 2000), and conditions that limit upregulation of OspC during the blood meal result in reduced spirochaete migration from the midgut to the salivary glands (de Silva et al., 1999; Gilmore and Piesman, 2000). B. burgdorferi s.s. can appear in the salivary glands of I. scapularis nymphs within 36 h of attachment but spirochaete numbers apparently increase dramatically after approximately 54 h (de Silva and Fikrig, 1995; Ohnishi et al., 2001). The time elapsed from tick attachment to spirochaete transmission may differ between vector species, between B. burgdorferi s.l. genospecies, and even between strains of the same genospecies. I. ricinus nymphs transmitted B. burgdorferi s.l. to approximately 50% of gerbil hosts within 17 h of attachment and all hosts developed viable infections after 48 h of nymphal feeding (Kahl et al., 1998). In contrast, only 11% of deer mice (Peromyscus maniculatus) fed on for 48 h by B. burgdorferi s.s.-infected I. pacificus nymphs developed viable infections, and approximately 78 h of nymphal attachment was required to transmit spirochaetes to 50% of the mice (Peavey and Lane, 1995). Early studies showed that I. scapularis nymphs transmitted B. burgdorferi s.s. strain JD1 to 7–14% of rodent hosts within 24–36 h of attachment, and 36–100% of the hosts developed viable infections after 48 h of nymphal feeding (e.g. Piesman et al., 1987; Shih and Spielman, 1993). Intriguingly, viable infections in white mice exposed to another B. burgdorferi s.s. strain (B31) occurred only for mice fed on by infected I. scapularis nymphs for at least 53 h, after which period spirochaete numbers increased dramatically in nymphal salivary glands and the majority of spirochaetes in the salivary glands failed to produce either OspA or OspC (Ohnishi et al., 2001). Finally, a recent comparative study recorded similar transmission rates of B. burgdorferi s.s. strains JD1 and B31 by I. scapularis nymphs, with 0–12% of white mice developing viable infections after 48 h of nymphal feeding and 56–71% after 72 h (des Vignes et al., 2001). In addition to the antigenic changes B. burgdorferi s.l. undergoes during the blood meal, the ability of these spirochaetes to infect the tick host successfully may be related to immunosuppressive factors present in tick saliva (e.g. Ribeiro and Mather, 1998; Valenzuela et al., 2000). For example, I. ricinus salivary gland extract inhibited killing of B. afzelii by mouse macrophages (Kuthejlová et al., 2001). Similarly, salivary components facilitated transmission of tick-borne encephalitis virus from tick to host (saliva-activated transmission) (Nuttall, 1999). The factor(s) ultimately determining vector competence are still unknown. Under experimental conditions, B. burgdorferi s.s. is readily detected in North American Dermacentor species during the first 2 weeks after detachment from an infective host but is usually absent after 3 weeks (e.g. Piesman and Sinsky, 1988; Dolan et al., 1997; Piesman and Happ, 1997). This phenomenon may be related to tick immune responses to B. burgdorferi s.l.; Johns et al. (2001) recorded strong borreliacidal activity in the haemolymph of D. variabilis, but not in that of I. scapularis. However, spirochaetal infection prevalences as high as 5.4% for A. americanum and 11.3% for Dermacentor reticulatus have been recorded in questing

Vectors of B. burgdorferi

99

ticks (Schulze et al., 1986; Kahl et al., 1992). This suggests that, under natural conditions, transstadial passage of B. burgdorferi s.l. in these ticks may be more common than indicated in laboratory studies (see Table 4.3). Vector incompetence could ultimately result from an inability of spirochaetes to escape the midgut during the blood meal, as suggested for D. occidentalis (Li and Lane, 1996), or spirochaete mortality caused by exposure to the haemocoel during the migration to the salivary glands (Johns et al., 2001).

Biology of the Primary LB Vectors As for all ixodid ticks, I. pacificus, I. persulcatus, I. ricinus and I. scapularis undergo four developmental stages: egg, larva, nymph and adult (Figs 4.2 and 4.3; Balashov, 1972; Oliver, 1989; Sonenshine, 1991). A single large blood meal is ingested in the larval, nymphal and adult female stages. After each blood meal three-host ticks such as Ixodes spp. detach from the host before they start moulting to the next stage or, in the case of adult females, laying eggs (Balashov, 1972). As in other Ixodes spp., adult male I. ricinus and I. persulcatus are considered facultative blood feeders and usually do not require a blood meal to fertilize females (Balashov, 1972). However, males may repeatedly take small blood meals and can remain on a host for weeks to months in search of females (Oliver, 1989). Mating can take place either in the vegetation or on the host while the female tick is feeding (e.g. Gray, 1987; Yuval and Spielman, 1990a). The mated female then lays a single egg batch; the average number of eggs produced typically exceeds 850 for I. pacificus and I. scapularis fed on laboratory rabbits, and 1000 for I. ricinus fed on sheep (Gray, 1981; Oliver et al., 1993a; Schoeler and Lane, 1993). The length of the developmental cycle depends on temperature conditions, seasonal timing of questing and host availability. I. scapularis ticks typically complete their developmental cycle in 2–3 years and I. pacificus, I. persulcatus and I. ricinus in 3 years, but ranges of 2–4 years for I. scapularis and 2–5/6 years for I. persulcatus and I. ricinus have been reported (e.g. Balashov, 1972; Yuval and Spielman, 1990b; Gray, 1991; Lindsay et al., 1998; Padgett and Lane, 2001). These ticks use an ambush strategy to contact hosts and rarely move more than a few metres while questing (e.g. Balashov, 1972; Gray, 1985; Loye and Lane, 1988; Falco and Fish, 1991). Hosts are detected by vibrations caused by animal movements, odours, body heat and shadows (Balashov, 1972; Sonenshine, 1993). The subadult stages feed on a wide range of hosts, including lizards, birds and mammals ranging in size from minuscule insectivores to large ungulates, whereas the adults are restricted to feeding on medium-sized to large mammals (Fig. 4.2; Anastos, 1957; Furman and Loomis, 1984; Jaenson et al., 1994; Keirans et al., 1996). The feeding process consists of an initial phase of slow feeding, characterized by ingestion of small amounts of blood but intensive cuticular growth in the tick midgut, followed by a phase of rapid ingestion of blood during the last 12–36 h of the blood meal; during the blood meal weight increases range from approximately 10–20-fold for larvae to more than 100-fold for females

100

L. Eisen and R.S. Lane

Fig. 4.2. Generalized developmental cycle for Ixodes pacificus, I. persulcatus, I. ricinus and I. scapularis. Tick hosts include various rodents (exemplified by a mouse and a squirrel), lizards, birds, insectivores (hedgehog) and mediumsized (hare) and large (deer) mammals. The importance of these hosts for each life stage is indicated by the scale of depiction. (Designed by B. Kaye.)

(Fig. 4.3; Balashov, 1972). Larval I. ricinus and I. persulcatus, for example, complete their feeding in 2–5 days, whereas the nymphs take 2–7 days, and the female ticks 6–11 days (Balashov, 1972). The feeding duration and success rate vary with choice of host, attachment site, host immune response and, in the case of exothermic lizard hosts, temperature (e.g. Balashov, 1972; Davidar et al., 1989; James and Oliver, 1990; Sonenshine, 1991; Lane, 1994). The long feeding period of ixodid ticks makes them vulnerable to host responses and, consequently, their saliva contains an array of components that serve as anticoagulants and mediate host inflammatory and immune responses (reviewed by Wikel, 1996; Bowman et al., 1997; Gillespie et al., 2000).

Vectors of B. burgdorferi

101

Fig. 4.3. Life stages of Ixodes ricinus. Left to right: larva, nymph, adult female, adult male; top row, unfed stages; bottom row, fed stages (female potentially fed; males do not engorge). (Photo by Gunilla Olsson, modified by Bernard Kaye.)

The density of questing ticks is related to climatic conditions and the abundance of vertebrate tick hosts. Local abundances of I. ricinus and I. scapularis ticks are typically positively correlated with those of deer (e.g. Gray et al., 1992; Wilson, 1998), which serve as important hosts for the adult tick stages (e.g. Tälleklint and Jaenson, 1994). Also, the density of I. scapularis nymphs in the northeastern United States was positively correlated with that of white-footed mice (Peromyscus leucopus) in the previous year (Ostfeld et al., 2001). However, less than 5% of a 2–3 year developmental cycle is spent in direct contact with hosts. Off-host survival is usually positively correlated with relative humidities surpassing 80% (MacLeod, 1934; Balashov, 1972; Stafford, 1994), when atmospheric water vapour can be actively taken up by the tick through the secretion and then ingestion of hygroscopic fluids (Lees, 1946; Kahl and Knülle, 1988). Active questing is also typically interrupted by periods when the ticks seek out humid microhabitats to regain lost body water (Lees and Milne, 1951; Loye and Lane, 1988). Because they are sensitive to desiccation, questing ticks tend to survive for longer periods of time in sheltered woodlands compared with more open, exposed habitats (e.g. Daniel et al., 1977; Bertrand and Wilson, 1996; Peavey and Lane, 1996; Lindsay et al., 1998), although in areas with wet and cool climates, such as the British Isles, tick survival can be similar in sheltered and open habitats (Gray, 1981). Seasonal patterns of questing activity can differ between tick species and with climatic conditions for the same tick species. For example, peak questing

102

L. Eisen and R.S. Lane

activity of the adult stage usually occurs during spring and early summer for I. ricinus and I. persulcatus, autumn for I. scapularis and winter for I. pacificus (e.g. Ai et al., 1990; Lane, 1990; Gray, 1991; Sonenshine, 1993; Korenberg, 1994). Depending, in part, on climatic conditions, the seasonal questing activity of I. ricinus can be continuous from spring to autumn, bimodal with peaks in spring and late summer or early autumn, or, at the southern limit of its geographical distribution, restricted to the winter months (Gray, 1991). Such seasonal patterns of questing activity are regulated primarily by the ticks entering long periods of diapause at various stages in their developmental cycles. This enables them to avoid questing during periods of the year with inclement climatic conditions, such as during the winter at high latitudes and during the hot summer months in southern Europe and the western USA. Diapause phenomena have been studied extensively for I. ricinus and I. persulcatus (reviewed by Belozerov, 1982; Gray, 1991), whereas fewer studies have focused on I. scapularis (Yuval and Spielman, 1990b; Daniels et al., 1996; Lindsay et al., 1998) and I. pacificus (Peavey and Lane, 1996; Padgett and Lane, 2001). On a large spatial scale, unfavourable climatic conditions limit the geographical distributions of the primary bridging vectors of LB spirochaetes. For example, these ticks cannot establish populations in areas with very short warm seasons; I. ricinus ticks are virtually absent in far northern Scandinavia and at high altitudes in central Europe (Daniel, 1993; Tälleklint and Jaenson, 1998). Likewise, I. pacificus ticks are limited to especially moist habitats in arid regions of the western USA, such as along rivers or at altitudes where snow melt provides sufficient moisture (Dennis et al., 1998).

Vector Potential for B. burgdorferi s.l. in Relation to Tick Ecology Tick vector potential for B. burgdorferi s.l. is influenced by factors intrinsic to the tick, e.g. tick questing behaviour, diapause phenomena and host preference, as well as phenomena extrinsic to the tick, such as climatic conditions, host abundance and behaviour, host immune response, host susceptibility to spirochaetes and genetic variation in infectivity of spirochaetes (Lane, 1994). Here we will give some examples of how the above-mentioned factors affect tick vector potential.

Seasonal patterns of questing activity Differences in the seasonality of questing by subadult ticks may influence vector potential for B. burgdorferi s.l. Peak infestation on rodents by I. scapularis nymphs in the northeastern USA and I. persulcatus nymphs in China precedes that of the larval stages by at least 1 month (e.g. Wilson and Spielman, 1985; Ai et al., 1991), a time interval sufficient for white-footed mice infected with B. burgdorferi s.l. by

Vectors of B. burgdorferi

103

feeding nymphs to have become infective to feeding ticks at the time of larval peak infestation (Donahue et al., 1987). These inverted patterns of seasonal activity of subadult ticks probably serve to intensify the enzootic transmission of B. burgdorferi s.l. (Spielman et al., 1985). No such clear temporal separation occurs for the peak abundances of larval and nymphal I. pacificus infesting lizards in California (Lane and Loye, 1989; Eisen et al., 2001) or for I. ricinus on rodents in Europe (e.g. Matuschka et al., 1990; Randolph et al., 1999), although peak activity by I. ricinus nymphs may precede that of the larvae in exposed habitats (Gray, 1991). Intriguingly, I. ricinus larvae can become infected with B. burgdorferi s.l. when feeding in close proximity to infected nymphs on uninfected hosts (co-feeding transmission; Gern and Rais, 1996). Although this phenomenon may increase the efficiency of enzootic transmission of B. burgdorferi s.l. in situations where peak larval and nymphal infestation coincide (Ogden et al., 1997), further studies are needed to assess the potential importance of co-feeding transmission in various enzootic cycles.

Role of different life stages as bridging vectors to humans The relative importance of different tick stages as bridging vectors of LB spirochaetes depends on how commonly they attach to humans, how frequently they are infected with spirochaetes and how likely they are to be detected and removed before spirochaetes are transmitted. The larval stage is of little importance as a bridging vector because larvae are not aggressive human biters and the prevalence of infection with B. burgdorferi s.l. is usually low in unfed larvae (reviewed by Lane, 1994; Mather and Ginsberg, 1994; Hubálek and Halouzka, 1998). The nymphal stage is considered the primary bridging vector for I. scapularis in the eastern USA and I. ricinus in Europe. In the eastern USA, most LB cases occur from June to August (Sonenshine, 1993; Falco et al., 1999; CDC, 2000), a time period with peak activity of questing I. scapularis nymphs but low activity of adult ticks (e.g. Sonenshine, 1993; Stafford and Magnarelli, 1993; Falco et al., 1999). Detected human bites by I. scapularis nymphs and adult females apparently are equally likely to result in LB (Sood et al., 1997). However, nymphal ticks typically feed on humans for a longer period of time than female ticks before being discovered (e.g. Yeh et al., 1995; Falco et al., 1996), and nymphs probably account for the vast majority of the up to 85% of bites that can go undetected in people developing erythema migrans (e.g. Walker, 1998). In Europe, the seasonal distribution of LB cases cannot be used to elucidate which tick stage is the primary bridging vector because the seasonal activity periods of nymphs and adults are similar (Gray, 1991). However, three European studies showed that nymphs accounted for as much as 54, 70 and 82%, respectively, of all detected human bites by I. ricinus ticks, as compared with 43, 22 and 3% for female ticks (Maiwald et al., 1998; Liebisch and Liebisch, 1999; Robertson et al., 2000).

104

L. Eisen and R.S. Lane

In contrast, I. persulcatus females are the primary bridging vector of LB spirochaetes to humans in far-eastern Europe and Asia because nymphs only rarely infest humans (Ai et al., 1990; Korenberg, 1994). Notably, the questing behaviour of the adult stage is more aggressive in I. persulcatus than in I. ricinus (Uspensky, 1993). Also, generalized infections (i.e. infections not restricted to the tick gut) were more common for B. afzelii or B. garinii in I. persulcatus adults (12.9% of infected ticks with generalized infections) than for B. burgdorferi s.s. in I. scapularis adults (2.4%) (Korenberg and Moskvitina, 1996). This finding suggests that I. persulcatus females may be more likely to transmit spirochaetes shortly after attachment. The relative importance of nymphal versus female ticks is harder to discern in western North America. The seasonal distribution of human LB cases in relation to the peak activity periods of I. pacificus nymphs (spring) and adults (late autumn to winter) indicated that the nymphal stage was the primary bridging vector in one rural area in northern California (Clover and Lane, 1995). However, there is probably considerable spatial variation in the importance of I. pacificus nymphs versus females as bridging vectors: humans are exposed to nymphs, but rarely to adults, in dense woodland with a ground cover dominated by leaf-litter, whereas adults, but very few nymphs, are collected by drag sampling in woodlands with emergent vegetation or open grasslands (e.g. Clover and Lane, 1995).

Complexity of enzootic B. burgdorferi s.l.–tick vector–vertebrate cycles Ironically, the most intensively studied enzootic maintenance cycle for B. burgdorferi s.l., occurring in the northeastern and north-central USA, may be the least complicated. In this region, the enzootic cycle appears to be dominated by a single genospecies (B. burgdorferi s.s.), a single tick vector (I. scapularis) and one primary vertebrate spirochaete reservoir (the white-footed mouse), although other vertebrate hosts for I. scapularis do contribute to the enzootic cycle and I. dentatus and lagomorphs maintain a Borrelia andersonii transmission cycle (see Piesman, Chapter 9). In contrast, at least two genospecies of B. burgdorferi s.l. (B. burgdorferi s.s., B. bissettii) are frequently isolated from ticks and/or rodents in the western and southeastern USA. The enzootic cycles in these regions also include multiple tick vectors of B. burgdorferi s.l. (I. angustus, I. pacificus and I. spinipalpis in the west and I. affinis, I. minor and I. scapularis in the southeast), and important vertebrate spirochaete reservoirs include woodrats/cotton rats and Peromyscus mice (see Piesman, Chapter 9). Finally, in the western USA I. pacificus subadults feed abundantly on lizards having a borreliacidal factor in their blood (e.g. Lane and Quistad, 1998), a phenomenon that probably serves to decrease the intensity of enzootic transmission of B. burgdorferi s.l. (zooprophylaxis; Spielman et al., 1985). In Europe, the enzootic cycle(s) maintained primarily by I. ricinus involves at

Vectors of B. burgdorferi

105

least five different genospecies of B. burgdorferi s.l. (B. burgdorferi s.s., B. afzelii, B. garinii, B. lusitaniae, B. valaisiana; see Gern and Huwait, Chapter 6). The roles of various vertebrates in the enzootic maintenance cycles of these spirochaetes need further study, although it appears that B. garinii and B. valaisiana may be associated primarily with birds and B. afzelii and B. burgdorferi s.s. with small mammals (see Kurtenbach et al., Chapter 5). In Asia, I. persulcatus and its vertebrate hosts maintain enzootic cycles of B. afzelii and B. garinii, and data on tick–spirochaete associations indicate the presence of several additional enzootic cycles including I. ovatus and B. japonica, I. tanuki and B. tanukii, I. turdus and B. turdi, and I. columnae, I. granulatus and I. nipponensis and B. valaisiana in eastern Asia (see Chapters 7 and 8). Notably, the geographical distributions of I. ricinus and I. persulcatus overlap in far-eastern Europe (see Chapter 6), which allows direct comparison of the roles of these species in enzootic transmission cycles and as bridging vectors to humans (Korenberg et al., 2001). The role of transovarially infected larvae in the maintenance of enzootic cycles is controversial (Randolph and Craine, 1995). Transovarially infected larvae may be more efficient enzootic vectors than nymphs in situations where extremely few nymphs feed on rodent reservoirs for B. burgdorferi s.l. and other enzootic tick vectors are rare or absent (e.g. Kurtenbach et al., 1995; Randolph and Craine, 1995). Spirochaete transmission by transovarially infected larvae was demonstrated by exposing white mice to groups of 30–50 I. ricinus larvae (Stanek et al., 1986), whereas a single B. burgdorferi s.l.-infected I. ricinus, I. pacificus or I. scapularis nymph can infect a susceptible rodent (Stanek et al., 1986; Donahue et al., 1987; Peavey and Lane, 1995). Further experimental studies are needed to evaluate the potential of transovarially infected larvae, as compared with transstadially infected nymphs, to transmit different B. burgdorferi s.l. genospecies to natural tick hosts.

Summary Vector competence for B. burgdorferi s.l. indicates that a tick species can acquire spirochaetes when feeding on an infective host, pass them transstadially and, subsequently, transmit the spirochaetes to a susceptible host while feeding. Experimental studies are required to demonstrate vector competence; observations of natural infections in ticks collected from hosts or vegetation do not provide definitive evidence of vector competence. Vector competence for B. burgdorferi s.l. has been confirmed for 12 species of ticks, including six of the 14 members of the I. ricinus complex (I. affinis, I. jellisoni, I. pacificus, I. persulcatus, I. ricinus, I. scapularis) and six other Ixodes species (I. angustus, I. dentatus, I. hexagonus, I. minor, I. muris, I. spinipalpis). In contrast, seven tick species evaluated for vector competence appear unable to transmit B. burgdorferi s.l. (A. americanum, D. andersoni, D. occidentalis, D. variabilis, I. cookei, I. holocyclus, I. ovatus). The potential of tick species, or their different life stages, to act as enzootic vectors or bridging vectors to humans is influenced by numerous factors, such as

106

L. Eisen and R.S. Lane

seasonality of questing activity, questing behaviour, host range and time elapsed from tick attachment to spirochaete transmission.

Acknowledgements We are indebted to Dr J.H. Oliver Jr, Georgia Southern University, for generously sharing unpublished data on vector competence studies. This work was supported by a grant from the National Institutes of Health (AI-22501) to R.S. Lane.

References Ai, C., Hu, R., Hyland, K.E., Wen, Y., Zhang, Y., Qiu, Q., Li, D., Liu, X., Shi, Z., Zhao, J. and Cheng, D. (1990) Epidemiological and aetiological evidence for transmission of Lyme disease by adult Ixodes persulcatus in an endemic area in China. International Journal of Epidemiology 19, 1061–1065. Ai, C., Qiu, G., Shi, Z., Wu, X., Liu, X. and Zhao, J. (1991) Host associations and seasonal abundance of immature Ixodes persulcatus (Acari: Ixodidae) in a Lyme-disease-endemic area in China. Experimental and Applied Acarology 12, 251–257. Alekseev, A.N., Dubinina, H.V., van de Pol, I. and Schouls, L.M. (2001) Identification of Ehrlichia spp. and Borrelia burgdorferi in Ixodes ticks in the Baltic Regions of Russia. Journal of Clinical Microbiology 39, 2237–2242. Anastos, G. (1957) The Ticks, or Ixodides, of the U.S.S.R. – a Review of the Literature. United States Department of Health, Education, and Welfare, Public Health Service Publication No. 458, 397 pp. Artsob, H. (1988) Powassan encephalitis. In: Monath, T.P. (ed.) The Arboviruses: Epidemiology and Ecology, Vol. IV. CRC Press, Boca Raton, Florida, pp. 29–49. Balashov, Y.S. (1972) Bloodsucking ticks (Ixodoidea) – vectors of diseases of man and animals. Miscellaneous Publications of

the Entomological Society of America 8, 163–376. Baranton, G., Marti Ras, N. and Postic, D. (1998) Molecular epidemiology of the aetiological agents of Lyme borreliosis. Wiener Klinische Wochenschrift 110, 850–855. Barker, I.K., Lindsay, L.R., Campbell, G.D., Surgeoner, G.A. and McEwen, S.A. (1993) The groundhog tick Ixodes cookei (Acari: Ixodidae): a poor potential vector of Lyme borreliosis. Journal of Wildlife Diseases 29, 416–422. Belozerov, V.N. (1982) Diapause and biological rhythms in ticks. In: Obenchain, F.D. and Galun, R. (eds) Physiology of Ticks. Pergamon Press, Oxford, UK, pp. 469–500. Bertrand, M.R. and Wilson, M.L. (1996) Microclimate-dependent survival of unfed adult Ixodes scapularis (Acari: Ixodidae) in nature: life cycle and study design implications. Journal of Medical Entomology 33, 619–627. Bowman, A.S., Coons, L.B., Needham, G.R. and Sauer, J.R. (1997) Tick saliva: recent advances and implications for vector competence. Medical and Veterinary Entomology 11, 277–285. Brown, R.N. and Lane, R.S. (1992) Lyme disease in California: a novel enzootic transmission cycle of Borrelia burgdorferi. Science 256, 1439–1442.

Vectors of B. burgdorferi

Burgdorfer, W., Barbour, A.G., Hayes, S.F., Benach, J.L., Grunwaldt, E. and Davis, J.P. (1982) Lyme disease – a tickborne spirochetosis? Science 216, 1317–1319. Burgdorfer, W., Barbour, A.G., Hayes, S.F., Péter, O. and Aeschlimann, A. (1983) Erythema chronicum migrans – a tickborne spirochetosis. Acta Tropica 40, 79–83. Burkot, T.R., Schneider, B.S., Pieniazek, N.J., Happ, C.M., Rutherford, J.S., Slemenda, S.B., Hoffmeister, E., Maupin, G.O. and Zeidner, N.S. (2000) Babesia microti and Borrelia bissettii transmission by Ixodes spinipalpis ticks among prairie voles, Microtus ochrogaster, in Colorado. Parasitology 121, 595–599. CDC (2000) Surveillance for Lyme disease – United States, 1992–1998. Morbidity and Mortality Weekly Report 49 (SS03), 1–11. Clover, J.R. and Lane, R.S. (1995) Evidence implicating nymphal Ixodes pacificus (Acari: Ixodidae) in the epidemiology of Lyme disease in California. American Journal of Tropical Medicine and Hygiene 53, 237–240. Costero, A. and Grayson, M.A. (1996) Experimental transmission of Powassan virus (Flaviviridae) by Ixodes scapularis ticks (Acari: Ixodidae). American Journal of Tropical Medicine and Hygiene 55, 536–546. Damrow, T., Freedman, H., Lane, R.S. and Preston, K.L. (1989) Is Ixodes (Ixodiopsis) angustus a vector of Lyme disease in Washington State? The Western Journal of Medicine 150, 580–582. Daniel, M. (1993) Influence of the microclimate on the vertical distribution of the tick Ixodes ricinus (L.) in Central Europe. Acarologia 34, 105–113. Daniel, M., Cerny, V., Dusbábek, F., Honzáková, E. and Olejnicek, J. (1977) Influence of microclimate on the life cycle of the common tick Ixodes ricinus

107

(L.) in an open area in comparison with forest habitats. Folia Parasitologica 24, 149–160. Daniels, T.J., Falco, R.C., Curran, K.L. and Fish, D. (1996) Timing of Ixodes scapularis (Acari: Ixodidae) oviposition and larval activity in southern New York. Journal of Medical Entomology 33, 140–147. Davidar, P., Wilson, M. and Ribeiro, J.M.C. (1989) Differential distribution of immature Ixodes dammini (Acari: Ixodidae) on rodent hosts. Journal of Parasitology 75, 898–904. Dennis, D.T., Nekomoto, T.S., Victor, J.C., Paul, W.S. and Piesman, J. (1998) Reported distribution of Ixodes scapularis and Ixodes pacificus (Acari: Ixodidae) in the United States. Journal of Medical Entomology 35, 629–638. Dolan, M.C., Maupin, G.O., Panella, N.A., Golde, W.T. and Piesman, J. (1997) Vector competence of Ixodes scapularis, I. spinipalpis, and Dermacentor andersoni (Acari: Ixodidae) in transmitting Borrelia burgdorferi, the etiologic agent of Lyme disease. Journal of Medical Entomology 34, 128–135. Dolan, M.C., Piesman, J., Mbow, M.L., Maupin, G.O., Péter, O., Brossard, M. and Golde, W.T. (1998) Vector competence of Ixodes scapularis and Ixodes ricinus (Acari: Ixodidae) for three genospecies of Borrelia burgdorferi. Journal of Medical Entomology 35, 465–470. Dolan, M.C., Lacombe, E.H. and Piesman, J. (2000) Vector competence of Ixodes muris (Acari: Ixodidae) for Borrelia burgdorferi. Journal of Medical Entomology 37, 766–768. Donahue, J.G., Piesman, J. and Spielman, A. (1987) Reservoir competence of white-footed mice for Lyme disease spirochetes. American Journal of Tropical Medicine and Hygiene 36, 92–96. Dumler, J.S. and Bakken, J.S. (1998) Human ehrlichiosis: newly recognized infections transmitted by ticks. Annual Review of Medicine 49, 201–213.

108

L. Eisen and R.S. Lane

Dumler, J.S., Barbet, A.F., Bekker, C.P.J., Dasch, G.A., Palmer, G.H., Ray, S.C., Rikihisa, Y. and Rurangirwa, F.R. (2001) Reorganization of genera in the families Rickettsiaceae and Anaplasmataceae in the order Rickettsiales: unification of some species of Ehrlichia with Anaplasma, Cowdria with Ehrlichia and Ehrlichia with Neorickettsia, descriptions of six new species combinations and designation of Ehrlichia equi and ‘HGE-agent’ as subjective synonyms of Ehrlichia phagocytophila. International Journal of Systematic and Evolutionary Microbiology 51, 2145–2165. Durden, L.A. and Keirans, J.E. (1996) Nymphs of the Genus Ixodes (Acari: Ixodidae) of the United States: Taxonomy, Identification Key, Distribution, Hosts, and Medical/ Veterinary Importance. The Entomological Society of America, Lanham, Maryland, 95 pp. Eisen, R.J., Eisen, L. and Lane, R.S. (2001) Prevalence and abundance of Ixodes pacificus immatures (Acari: Ixodidae) infesting western fence lizards (Sceloporus occidentalis) in northern California: temporal trends and environmental correlates. Journal of Parasitology 87, 1301–1307. Falco, R.C. and Fish, D. (1991) Horizontal movement of adult Ixodes dammini (Acari: Ixodidae) attracted to CO2-baited traps. Journal of Medical Entomology 28, 726–729. Falco, R.C., Fish, D. and Piesman, J. (1996) Duration of tick bites in a Lyme diseaseendemic area. American Journal of Epidemiology 143, 187–192. Falco, R.C., McKenna, D.F., Daniels, T.J., Nadelman, R.B., Nowakowski, J., Fish, D. and Wormser, G.P. (1999) Temporal relation between Ixodes scapularis abundance and risk for Lyme disease associated with erythema migrans. American Journal of Epidemiology 149, 771–776. Filippova, N.A. (1999) Systematic relationships of the Ixodes ricinus complex in the

palearctic faunal region. In: Needham, G.R., Mitchell, R., Horn, D.J and Welbourn, W.C. (eds) Acarology IX: Vol. 2, Symposia. Ohio Biological Survey, Columbus, Ohio, pp. 355–361. Fukunaga, M., Takahashi, Y., Tsuruta, Y., Matsushita, O., Ralph, D., McClelland, M. and Nakao, M. (1995) Genetic and phenotypic analysis of Borrelia miyamotoi sp. nov., isolated from the ixodid tick Ixodes persulcatus, the vector for Lyme disease in Japan. International Journal of Systematic Bacteriology 45, 804–810. Fukunaga, M., Hamase, A., Okada, K., Inoue, H., Tsuruta, Y., Miyamoto, K. and Nakao, M. (1996) Characterization of spirochetes isolated from ticks (Ixodes tanuki, Ixodes turdus, and Ixodes columnae) and comparison of the sequences with those of Borrelia burgdorferi sensu lato strains. Applied and Environmental Microbiology 62, 2338–2344. Furman, D.P. and Loomis, E.C. (1984) The ticks of California (Acari: Ixodida). Bulletin of the California Insect Survey 25, 1–239. Gern, L. and Rais, O. (1996) Efficient transmission of Borrelia burgdorferi between cofeeding Ixodes ricinus ticks (Acari: Ixodidae). Journal of Medical Entomology 33, 189–192. Gern, L., Zhu, Z. and Aeschlimann, A. (1990) Development of Borrelia burgdorferi in Ixodes ricinus females during blood feeding. Annales de Parasitologie Humaine et Comparée 65, 89–93. Gern, L., Toutoungi, L.N., Hu, C.M. and Aeschlimann, A. (1991) Ixodes (Pholeoixodes) hexagonus, an efficient vector of Borrelia burgdorferi in the laboratory. Medical and Veterinary Entomology 5, 431–435. Gillespie, R.D., Mbow, M.L. and Titus, R.G. (2000) The immunomodulatory factors of bloodfeeding arthropod saliva. Parasite Immunology 22, 319–331. Gilmore, R.D. Jr and Piesman, J. (2000) Inhibition of Borrelia burgdorferi migration from the midgut to the salivary glands

Vectors of B. burgdorferi

following feeding by ticks on OspCimmunized mice. Infection and Immunity 68, 411–414. Gray, J.S. (1981) The fecundity of Ixodes ricinus (L.) (Acarina: Ixodidae) and the mortality of its developmental stages under field conditions. Bulletin of Entomological Research 71, 533–542. Gray, J.S. (1985) A carbon dioxide trap for prolonged sampling of Ixodes ricinus L. populations. Experimental and Applied Acarology 1, 35–44. Gray, J.S. (1987) Mating and behavioural diapause in Ixodes ricinus L. Experimental and Applied Acarology 3, 61–71. Gray, J.S. (1991) The development and seasonal activity of the tick Ixodes ricinus: a vector of Lyme borreliosis. Review of Medical and Veterinary Entomology 79, 323–333. Gray, J.S., Kahl, O., Janetzki, C. and Stein, J. (1992) Studies on the ecology of Lyme disease in a deer forest in County Galway, Ireland. Journal of Medical Entomology 29, 915–920. Hu, C.M., Wilske, B., Fingerle, V., Lobet, Y. and Gern, L. (2001) Transmission of Borrelia garinii OspA serotype 4 to BALB/c mice by Ixodes ricicnus ticks collected in the field. Journal of Clinical Microbiology 39, 1169–1171. Hubálek, Z. and Halouzka, J. (1998) Prevalence rates of Borrelia burgdorferi sensu lato in host-seeking Ixodes ricinus ticks in Europe. Parasitology Research 84, 167–172. Hubálek, Z., Halouzka, J. and Juricová, Z. (1998) Investigation of haematophagous arthropods for borreliae – summarized data, 1988–1996. Folia Parasitologica 45, 67–72. Jaenson, T.G.T., Tälleklint, L., Lundqvist, L., Olsen, B., Chirico, J. and Mejlon, H.A. (1994) Geographical distribution, host associations, and vector roles of ticks (Acari: Ixodidae, Argasidae) in Sweden. Journal of Medical Entomology 31, 240–256. James, A.M. and Oliver, J.H. Jr (1990)

109

Feeding and host preference of immature Ixodes dammini, I. scapularis, and I. pacificus (Acari: Ixodidae). Journal of Medical Entomology 27, 324–330. Johns, R., Ohnishi, J., Broadwater, A., Sonenshine, D.E., de Silva, A.M. and Hynes, W.L. (2001) Contrasts in tick innate immune responses to Borrelia burgdorferi challenge: immunotolerance in Ixodes scapularis versus immunocompetence in Dermacentor variabilis (Acari: Ixodidae). Journal of Medical Entomology 38, 99–107. Kahl, O. and Knülle, W. (1988) Water vapour uptake from subsaturated atmospheres by engorged immature ixodid ticks. Experimental and Applied Acarology 4, 73–83. Kahl, O., Janetzki, C., Gray, J.S., Stein, J. and Bauch, R.J. (1992) Tick infection rates with Borrelia: Ixodes ricinus versus Haemaphysalis concinna and Dermacentor reticulatus in two locations in eastern Germany. Medical and Veterinary Entomology 6, 363–366. Kahl, O., Janetzki-Mittman, C., Gray, J.S., Jonas, R., Stein, J. and de Boer, R. (1998) Risk of infection with Borrelia burgdorferi sensu lato for a host in relation to the duration of nymphal Ixodes ricinus feeding and the method of tick removal. Zentralblatt für Bakteriologie 287, 41–52. Keirans, J.E., Hutcheson, H.J., Durden, L.A. and Klompen, J.S.H. (1996) Ixodes (Ixodes) scapularis (Acari: Ixodidae): redescription of all active stages, distribution, hosts, geographical variation, and medical and veterinary importance. Journal of Medical Entomology 33, 297–318. Keirans, J.E., Needham, G.R. and Oliver, J.H. Jr (1999) The Ixodes ricinus complex worldwide: diagnosis of the species in the complex, hosts and distribution. In: Needham, G.R., Mitchell, R., Horn, D.J and Welbourn, W.C. (eds) Acarology IX: Vol. 2, Symposia. Ohio Biological Survey, Columbus, Ohio, pp. 341–347.

110

L. Eisen and R.S. Lane

Kjemtrup, A.M. and Conrad, P.A. (2000) Human babesiosis: an emerging tickborne disease. International Journal for Parasitology 30, 1323–1337. Korenberg, E.I. (1994) Comparative ecology and epidemiology of Lyme disease and tick-borne encephalitis in the former Soviet Union. Parasitology Today 10, 157–160. Korenberg, E.I. and Moskvitina, G.G. (1996) Interrelationships between different Borrelia genospecies and their principal vectors. Journal of Vector Ecology 21, 178–185. Korenberg, E.I., Kovalevskii, Y.V., Levin, M.L. and Shchyogoleva, T.V. (2001) The prevalence of Borrelia burgdorferi sensu lato in Ixodes persulcatus and I. ricinus ticks in the zone of their sympatry. Folia Parasitologica 48, 63–68. Kurtenbach, K., Kampen, H., Dizij, A., Arndt, S., Seitz, H.M., Schaible, U.E. and Simon, M.M. (1995) Infestation of rodents with larval Ixodes ricinus (Acari: Ixodidae) is an important factor in the transmission cycle of Borrelia burgdorferi s.l. in German woodlands. Journal of Medical Entomology 32, 807–817. Kuthejlová, M., Kopecky, J., Stepanová, G. and Macela, A. (2001) Tick salivary gland extract inhibits killing of Borrelia afzelii spirochetes by mouse macrophages. Infection and Immunity 69, 575–578. Lane, R.S. (1990) Seasonal activity of two human-biting ticks. California Agriculture 44, 23–25. Lane, R.S. (1994) Competence of ticks as vectors of microbial agents with an emphasis on Borrelia burgdorferi. In: Sonenshine, D.E. and Mather, T.N. (eds) Ecological Dynamics of Tick-borne Zoonoses. Oxford University Press, Oxford, UK, pp. 45–67. Lane, R.S. and Loye, J.E. (1989) Lyme disease in California: interrelationship of Ixodes pacificus (Acari: Ixodidae), the western fence lizard (Sceloporus occidentalis), and Borrelia burgdorferi. Journal of Medical Entomology 26, 272–278.

Lane, R.S. and Quistad, G.B. (1998) Borreliacidal factor in the blood of the western fence lizard (Sceloporus occidentalis). Journal of Parasitology 84, 29–34. Lane, R.S., Brown, R.N., Piesman, J. and Peavey, C.A. (1994) Vector competence of Ixodes pacificus and Dermacentor occidentalis (Acari: Ixodidae) for various isolates of Lyme disease spirochetes. Journal of Medical Entomology 31, 417–424. Lane, R.S., Peavey, C.A., Padgett, K.A. and Hendson, M. (1999) Life history of Ixodes (Ixodes) jellisoni (Acari: Ixodidae) and its vector competence for Borrelia burgdorferi sensu lato. Journal of Medical Entomology 36, 329–340. Lees, A.D. (1946) The water balance in Ixodes ricinus L. and certain other species of ticks. Parasitology 37, 1–20. Lees, A.D. and Milne, A. (1951) The seasonal and diurnal activities of individual sheep ticks (Ixodes ricinus L.). Parasitology 41, 189–208. Leuba-Garcia, S., Martinez, R. and Gern, L. (1998) Expression of outer surface proteins A and C of Borrelia afzelii in Ixodes ricinus ticks and in the skin of mice. Zentralblatt für Bakteriologie 287, 475–484. Li, X. and Lane, R.S. (1996) Vector competence of ixodid ticks (Acari) for Borrelia burgdorferi as determined with a capillary-feeding technique. Journal of Spirochetal and Tick-borne Diseases 3, 116–123. Liebisch, A. and Liebisch, G. (1999) Hard ticks (Ixodidae) biting humans in Germany and their infection with Borrelia burgdorferi. In: Needham, G.R., Mitchell, R., Horn, D.J. and Welbourn, W.C. (eds) Acarology IX: Vol. 2, Symposia. Ohio Biological Survey, Columbus, Ohio, pp. 465–468. Lindsay, L.R., Barker, I.K., Surgeoner, G.A., McEwen, S.A., Gillespie, T.J. and Addison, E.M. (1998) Survival and development of the different life stages of Ixodes scapularis (Acari: Ixodidae) held

Vectors of B. burgdorferi

within four habitats on Long Point, Ontario, Canada. Journal of Medical Entomology 35, 189–199. Loye, J.E. and Lane, R.S. (1988) Questing behavior of Ixodes pacificus (Acari: Ixodidae) in relation to meteorological and seasonal factors. Journal of Medical Entomology 25, 391–398. MacLeod, J. (1934) Ixodes ricinus in relation to its physical environment: the influence of climate on development. Parasitology 26, 282–305. Magnarelli, L.A. and Anderson, J.F. (1988) Ticks and biting insects infected with the etiologic agent of Lyme disease, Borrelia burgdorferi. Journal of Clinical Microbiology 26, 1482–1486. Maiwald, M., Oehme, R., March, O., Petney, T.N., Kimmig, P., Naser, K., Zappe, H.A., Hassler, D. and von Knebel Doeberitz, M. (1998) Transmission risk of Borrelia burgdorferi sensu lato from Ixodes ricinus ticks to humans in southwest Germany. Epidemiology and Infection 121, 103–108. Masuzawa, T., Fukui, T., Miyake, M., Oh, H.-B., Cho, M.-K., Chang, W.-H., Imai, Y. and Yanagihara, Y. (1999) Determination of members of a Borrelia afzelii-related group isolated from Ixodes nipponensis in Korea as Borrelia valaisiana. International Journal of Systematic Bacteriology 49, 1409–1415. Mather, T.N. and Ginsberg, H.S. (1994) Vector–host–pathogen relationships: transmission dynamics of tick-borne infections. In: Sonenshine, D.E. and Mather, T.N. (eds) Ecological Dynamics of Tick-borne Zoonoses. Oxford University Press, Oxford, UK, pp. 68–90. Mather, T.N. and Mather, M.E. (1990) Intrinsic competence of three ixodid ticks (Acari) as vectors of the Lyme disease spirochete. Journal of Medical Entomology 27, 646–650. Matuschka, F.-R., Lange, R., Spielman, A., Richter, D. and Fischer, P. (1990) Subadult Ixodes ricinus (Acari: Ixodidae) on rodents in Berlin, West Germany.

111

Journal of Medical Entomology 27, 385–390. Mukolwe, S.W., Kocan, A.A., Barker, R.W., Kocan, K.M. and Murphy, G.L. (1992) Attempted transmission of Borrelia burgdorferi (Spirochaetales: Spirochaetaceae) (JD1 strain) by Ixodes scapularis (Acari: Ixodidae), Dermacentor variabilis, and Amblyomma americanum. Journal of Medical Entomology 29, 673–677. Nakao, M. and Miyamoto, K. (1994) Susceptibility of Ixodes persulcatus and I. ovatus (Acari: Ixodidae) to Lyme disease spirochetes isolated from humans in Japan. Journal of Medical Entomology 31, 467–473. Nakao, M. and Sato, Y. (1996) Refeeding activity of immature ticks of Ixodes persulcatus and transmission of Lyme disease spirochete by partially fed larvae. Journal of Parasitology 84, 669–672. Norris, D.E., Klompen, J.S.H., Keirans, J.E., Lane, R.S., Piesman, J. and Black, W.C. IV (1997) Taxonomic status of Ixodes neotomae and I. spinipalpis (Acari: Ixodidae) based on mitochondrial DNA evidence. Journal of Medical Entomology 34, 696–703. Nuttall, P.A. (1999) Pathogen–tick–host interactions: Borrelia burgdorferi and TBE virus. Zentralblatt für Bakteriologie 289, 492–505. Nuttall, P.A. and Labuda, M. (1994) Tickborne encephalitis subgroup. In: Sonenshine, D.E. and Mather, T.N. (eds) Ecological Dynamics of Tick-borne Zoonoses. Oxford University Press, Oxford, pp. 351–391. Ogden, N.H., Nuttall, P.A. and Randolph, S.E. (1997) Natural Lyme disease cycles maintained via sheep by cofeeding ticks. Parasitology 115, 591–599. Ohnishi, J., Piesman, J. and de Silva, A.M. (2001) Antigenic and genetic heterogeneity of Borrelia burgdorferi populations transmitted by ticks. Proceedings of the National Academy of Sciences, USA 98, 670–675.

112

L. Eisen and R.S. Lane

Oliver, J.H. Jr (1989) Biology and systematics of ticks (Acari: Ixodida). Annual Review of Ecology and Systematics 20, 397–430. Oliver, J.H. Jr (1996) Lyme borreliosis in the southern United States: a review. Journal of Parasitology 82, 926–935. Oliver, J.H. Jr, Owsley, M.R., Hutcheson, H.J., James, A.M., Chen, C., Irby, W.S., Dotson, E.M. and McLain, D.K. (1993a) Conspecificity of the ticks Ixodes scapularis and I. dammini (Acari: Ixodidae). Journal of Medical Entomology 30, 54–63. Oliver, J.H. Jr, Chandler, F.W. Jr, Luttrell, M.P., James, A.M., Stallknecht, D.E., McGuire, B.S., Hutcheson, H.J., Cummins, G.A. and Lane, R.S. (1993b) Isolation and transmission of the Lyme disease spirochete from the southeastern United States. Proceedings of the National Academy of Sciences, USA 90, 7371–7375. Olsen, B., Duffy, D.C., Jaenson, T.G.T., Gylfe, Å., Bonnedahl, J. and Bergström, S. (1995) Transhemispheric exchange of Lyme disease spirochetes by seabirds. Journal of Clinical Microbiology 33, 3270–3274. Ostfeld, R.S., Schauber, E.M., Canham, C.D., Keesing, F., Jones, C.G. and Wolff, J.O. (2001) Effects of acorn production and mouse abundance on abundance and Borrelia burgdorferi infection prevalence of nymphal Ixodes scapularis ticks. Vector Borne and Zoonotic Diseases 1, 55–63. Padgett, K.A. and Lane, R.S. (2001) Life cycle of Ixodes pacificus (Acari: Ixodidae): timing of developmental processes under field and laboratory conditions. Journal of Medical Entomology 38, 684–693. Pal, U., de Silva, A.M., Montgomery, R.R., Fish, D., Anguita, J., Anderson, J.F., Lobet, Y. and Fikrig, E. (2000) Attachment of Borrelia burgdorferi within Ixodes scapularis mediated by outer surface protein A. Journal of Clinical Investigation 106, 561–569.

Peavey, C.A. and Lane, R.S. (1995) Transmission of Borrelia burgdorferi by Ixodes pacificus nymphs and reservoir competence of deer mice (Peromyscus maniculatus) infected by tick-bite. Journal of Parasitology 81, 175–178. Peavey, C.A. and Lane, R.S. (1996) Field and laboratory studies on the timing of oviposition and hatching of the western black-legged tick, Ixodes pacificus (Acari: Ixodidae). Experimental and Applied Acarology 20, 695–711. Peavey, C.A., Lane, R.S. and Damrow, T. (2000) Vector competence of Ixodes angustus (Acari: Ixodidae) for Borrelia burgdorferi sensu stricto. Experimental and Applied Acarology 24, 77–84. Piesman, J. and Happ, C.M. (1997) Ability of the Lyme disease spirochete Borrelia burgdorferi to infect rodents and three species of human-biting ticks (Blacklegged tick, American dog tick, Lone star tick) (Acari: Ixodidae). Journal of Medical Entomology 34, 451–456. Piesman, J. and Sinsky, R.J. (1988) Ability of Ixodes scapularis, Dermacentor variabilis, and Amblyomma americanum (Acari: Ixodidae) to acquire, maintain, and transmit Lyme disease spirochetes (Borrelia burgdorferi). Journal of Medical Entomology 25, 336–339. Piesman, J. and Stone, B.F. (1991) Vector competence of the Australian paralysis tick, Ixodes holocyclus, for the Lyme disease spirochete Borrelia burgdorferi. International Journal for Parasitology 21, 109–111. Piesman, J., Mather, T.N., Sinsky, R.J. and Spielman, A. (1987) Duration of tick attachment and Borrelia burgdorferi transmission. Journal of Clinical Microbiology 25, 557–558. Piesman, J., Oliver, J.R. and Sinsky, R.J. (1990) Growth kinetics of the Lyme disease spirochete (Borrelia burgdorferi) in vector ticks (Ixodes dammini). American Journal of Tropical Medicine and Hygiene 42, 352–357. Postic, D., Korenberg, E., Gorelova, N., Kovalevski, Y.V., Bellenger, E. and

Vectors of B. burgdorferi

Baranton, G. (1997) Borrelia burgdorferi sensu lato in Russia and neighbouring countries: high incidence of mixed isolates. Research in Microbiology 148, 691–702. Randolph, S.E. and Craine, N.G. (1995) General framework for comparative quantitative studies on transmission of tick-borne diseases using Lyme borreliosis in Europe as an example. Journal of Medical Entomology 32, 765–777. Randolph, S.E. and Nuttall, P.A. (1994) Nearly right or precisely wrong? Natural versus laboratory studies of vector-borne diseases. Parasitology Today 10, 458–462. Randolph, S.E., Miklisová, D., Lysy, J., Rogers, D.J. and Labuda, M. (1999) Incidence from coincidence: patterns of tick infestations on rodents facilitate transmission of tick-borne encephalitis virus. Parasitology 118, 177–186. Ribeiro, J.M.C. and Mather, T.N. (1998) Ixodes scapularis: salivary kininase activity is a metallo dipeptidyl carboxypeptidase. Experimental Parasitology 89, 213–221. Ribeiro, J.M.C., Mather, T.N., Piesman, J. and Spielman, A. (1987) Dissemination and salivary delivery of Lyme disease spirochetes in vector ticks (Acari: Ixodidae). Journal of Medical Entomology 24, 201–205. Robertson, J.N., Gray, J.S. and Stewart, P. (2000) Tick bite and Lyme borreliosis risk at a recreational site in England. European Journal of Epidemiology 16, 647–652. Ryder, J.W., Pinger, R.R. and Glancy, T. (1992) Inability of Ixodes cookei and Amblyomma americanum nymphs (Acari: Ixodidae) to transmit Borrelia burgdorferi. Journal of Medical Entomology 29, 525–530. Sanders, F.H. Jr and Oliver, J.H. Jr (1995) Evaluation of Ixodes scapularis, Amblyomma americanum, and Dermacentor variabilis (Acari: Ixodidae) from Georgia as vectors of a Florida strain of the

113

Lyme disease spirochete, Borrelia burgdorferi. Journal of Medical Entomology 32, 402–406. Sato, Y. and Nakao, M. (1997) Transmission of the Lyme disease spirochete, Borrelia garinii, between infected and uninfected immature Ixodes persulcatus during cofeeding on mice. Journal of Parasitology 83, 547–550. Schoeler, G.B. and Lane, R.S. (1993) Efficiency of transovarial transmission of the Lyme disease spirochete, Borrelia burgdorferi, in the western blacklegged tick, Ixodes pacificus (Acari: Ixodidae). Journal of Medical Entomology 30, 80–86. Schulze, T.L., Lakat, M.F., Parkin, W.E., Shisler, J.K., Charette, D.J. and Bosler, E.M. (1986) Comparison of rates of infection by the Lyme disease spirochete in selected populations of Ixodes dammini and Amblyomma americanum (Acari: Ixodidae). Zentralblatt für Bakteriologie, Mikrobiologie, und Hygiene Series A 263, 72–78. Schwan, T.G., Piesman, J., Golde, W.T., Dolan, M.C. and Rosa, P.A. (1995) Induction of an outer surface protein on Borrelia burgdorferi during tick feeding. Proceedings of the National Academy of Sciences, USA 92, 2909–2913. Scoles, G.A., Papero, M., Beati, L. and Fish, D. (2001) A relapsing fever group spirochete transmitted by Ixodes scapularis ticks. Vector Borne and Zoonotic Diseases 1, 21–34. Shih, C.-M. and Spielman, A. (1993) Accelerated transmission of Lyme disease spirochetes by partially fed vector ticks. Journal of Clinical Microbiology 31, 2878–2881. de Silva, A.M. and Fikrig, E. (1995) Growth and migration of Borrelia burgdorferi in Ixodes ticks during blood feeding. American Journal of Tropical Medicine and Hygiene 53, 397–404. de Silva, A.M., Telford, S.R. III, Brunet, L.R., Barthold, S.W. and Fikrig, E. (1996) Borrelia burgdorferi OspA is an

114

L. Eisen and R.S. Lane

arthropod-specific transmission-blocking Lyme disease vaccine. Journal of Experimental Medicine 183, 271–275. de Silva, A.M., Zeidner, N.S., Zhang, Y., Dolan, M.C., Piesman, J. and Fikrig, E. (1999) Influence of outer surface protein A antibody on Borrelia burgdorferi within feeding ticks. Infection and Immunity 67, 30–35. Sonenshine, D.E. (1991) Biology of Ticks, Vol. 1. Oxford University Press, Oxford, UK, 447 pp. Sonenshine, D.E. (1993) Biology of Ticks, Vol. 2. Oxford University Press, Oxford, UK, 465 pp. Sood, S.K., Salzman, M.B., Johnson, B.J.B., Happ, C.M., Feig, K., Carmody, L., Rubin, L.G., Hilton, E. and Piesman, J. (1997) Duration of tick attachment as a predictor of the risk of Lyme disease in an area in which Lyme disease is endemic. Journal of Infectious Diseases 175, 996–999. Spielman, A., Wilson, M.L., Levine, J.F. and Piesman, J. (1985) Ecology of Ixodes dammini-borne human babesiosis and Lyme disease. Annual Review of Entomology 30, 439–460. Stafford, K.C. III (1994) Survival of immature Ixodes scapularis (Acari: Ixodidae) at different relative humidities. Journal of Medical Entomology 31, 310–314. Stafford, K.C. III and Magnarelli, L.A. (1993) Spatial and temporal patterns of Ixodes scapularis (Acari: Ixodidae) in southeastern Connecticut. Journal of Medical Entomology 30, 762–771. Stanek, G., Burger, I., Hirschl, A., Wewalka, G. and Radda, A. (1986) Borrelia transfer by ticks during their life cycle. Zentralblatt für Bakteriologie, Mikrobiologie, und Hygiene Series A 263, 29–33. Tälleklint, L. and Jaenson, T.G.T. (1994) Transmission of Borrelia burgdorferi s.l. from mammal reservoirs to the primary vector of Lyme borreliosis, Ixodes ricinus (Acari: Ixodidae), in Sweden. Journal of Medical Entomology 31, 880–886. Tälleklint, L. and Jaenson, T.G.T. (1998)

Increasing geographical distribution and density of Ixodes ricinus (Acari: Ixodidae) in central and northern Sweden. Journal of Medical Entomology 35, 521–526. Telford, S.R. III and Spielman, A. (1989a) Competence of a rabbit-feeding Ixodes (Acari: Ixodidae) as a vector of the Lyme disease spirochete. Journal of Medical Entomology 26, 118–121. Telford, S.R. III and Spielman, A. (1989b) Enzootic transmission of the agent of Lyme disease in rabbits. American Journal of Tropical Medicine and Hygiene 41, 482–490. Telford, S.R. III, Armstrong, P.M., Katavolos, P., Foppa, I., Olmeda Garcia, A.S., Wilson, M.L. and Spielman, A. (1997) A new tickborne encephalitis-like virus infecting New England deer ticks, Ixodes dammini. Emerging Infectious Diseases 3, 165–170. Uspensky, I. (1993) Ability of successful attack in two species of ixodid ticks (Acari: Ixodidae) as a manifestation of their aggressiveness. Experimental and Applied Acarology 17, 673–683. Valenzuela, J.G., Charlab, R., Mather, T.N. and Ribeiro, J.M.C. (2000) Purification, cloning, and expression of a novel salivary anticomplement protein from the tick, Ixodes scapularis. Journal of Biological Chemistry 275, 18717–18723. des Vignes, F., Piesman, J., Heffernan, R., Schulze, T.L., Stafford, K.C. III and Fish, D. (2001) Effect of tick removal on transmission of Borrelia burgdorferi and Ehrlichia phagocytophila by Ixodes scapularis nymphs. Journal of Infectious Diseases 183, 773–778. Walker, D.H. (1998) Tick-transmitted infectious diseases in the United States. Annual Review of Public Health 19, 237–269. Wang, G., van Dam, A.P., Schwartz, I. and Dankert, J. (1999) Molecular typing of Borrelia burgdorferi sensu lato: taxonomic, epidemiological, and clinical implica-

Vectors of B. burgdorferi

tions. Clinical Microbiology Reviews 12, 633–653. Wikel, S.K. (1996) Host immunity to ticks. Annual Review of Entomology 41, 1–22. Wilson, M.L. (1998) Distribution and abundance of Ixodes scapularis (Acari: Ixodidae) in North America: ecological processes and spatial analysis. Journal of Medical Entomology 35, 446–457. Wilson, M.L. and Spielman, A. (1985) Seasonal activity of immature Ixodes dammini (Acari: Ixodidae). Journal of Medical Entomology 22, 408–414. Woldehiwet, Z. (1983) Tick-borne fever: a review. Veterinary Research Communications 6, 163–175.

115

Yeh, M.-T., Bak, J.M., Hu, R., Nicholson, M.C., Kelly, C. and Mather, T.N. (1995) Determining the duration of Ixodes scapularis (Acari: Ixodidae) attachment to tick-bite victims. Journal of Medical Entomology 32, 853–858. Yuval, B. and Spielman, A. (1990a) Sperm precedence in the deer tick Ixodes dammini. Physiological Entomology 15, 123–128. Yuval, B. and Spielman, A. (1990b) Duration and regulation of the developmental cycle of Ixodes dammini (Acari: Ixodidae). Journal of Medical Entomology 27, 196–201.`

Borrelia burgdorferi sensu lato in the Vertebrate Host

5

Klaus Kurtenbach,1 Stefanie M. Schäfer,1,2 Simona de Michelis,1 Susanne Etti1,2 and Henna-Sisko Sewell1,2 1

Department of Infectious Disease Epidemiology, Imperial College of Science, Technology and Medicine, St Mary’s Campus, Norfolk Place, London W2 1PG, UK; 2Department of Zoology, University of Oxford, Oxford, UK

Introduction Borrelia burgdorferi sensu lato (s.l.) is a bacterial species complex, at present comprising 11 delineated and named genospecies (Postic et al., 1994, 1998; Le Fleche et al., 1997; Masuzawa et al., 2001). In addition, several as yet unnamed genomic groups contribute to the increasingly recognized diversity of B. burgdorferi s.l. Like all the other members of the genus Borrelia, the life style of B. burgdorferi s.l. is obligate parasitic, with no free-living stages known. B. burgdorferi s.l. is maintained in nature by complex zoonotic transmission cycles involving a large variety of mammalian and avian hosts and hard ticks of the genus Ixodes as vectors. Humans are a dead-end host. Larvae, nymphs and female ticks each feed once on a host for several consecutive days. Once infected by an infectious vector tick, reservoir hosts develop a prolonged or even persistent spirochaete infection, which can efficiently be passed on to uninfected ticks (Piesman et al., 1987, 1991; Piesman, 1988, 1993a). Natural hosts do not appear to develop clinical disease (Donahue et al., 1987; Gern et al., 1994; Kurtenbach et al., 1994, 1995); however, minor disease symptoms may escape detection as they are possibly confounded by other factors such as increased predation. Rodents are considered to be the principal reservoir hosts (Donahue et al., 1987), but the role avian hosts play in the ecology of Lyme borreliosis (LB) is increasingly gaining recognition (Humair et al., 1998; Kurtenbach et al., 1998a,b; Gylfe et al., 2000; Richter et al., 2000). A variety of laboratory animals are known to develop disease (Schaible et al., 1989, 1994; Barthold, 1995; Barthold et al., 1999). In this chapter the major principles governing transmission, dissemination, tissue invasion, survival and pathogenesis of B. burgdorferi s.l. in the vertebrate host are outlined. In addition, special emphasis is placed on the diversity of these © 2002 CAB International. Lyme Borreliosis: Biology, Epidemiology and Control (eds J. Gray, O. Kahl, R.S. Lane and G. Stanek)

117

118

K. Kurtenbach et al.

processes, as differences in genetic background, physiology, immunology, behaviour and ecology of vertebrate hosts and Borrelia spp. genotypes are important elements in the eco-epidemiology and pathogenesis of LB.

Current Concepts of Transmission, Dissemination, Persistence and Pathogenesis of B. burgdorferi s.l. in the Vertebrate Host Transmission of B. burgdorferi s.l. from the tick to the host Transmission of B. burgdorferi s.l. from the tick to the host is not a mechanical event, but rather a complex biological process that has only recently been recognized (Fikrig et al., 1992; de Silva and Fikrig, 1997; de Silva et al., 1999; Ohnishi et al., 2001). One of the major features that has emerged is that the life cycle of B. burgdorferi s.l. is highly adapted and is analogous to transmission cycles of eukaryotic vector-borne parasites, for example Plasmodium spp. (de Silva and Fikrig, 1997). Most members of the Ixodes ricinus/persulcatus species complex are known to be vector-competent for B. burgdorferi s.l. According to the current model of transmission, B. burgdorferi s.l. is normally confined to the midgut of unfed infected ticks (Lebet and Gern, 1994; de Silva and Fikrig, 1997; de Silva et al., 1999; Ohnishi et al., 2001). It is now well documented that B. burgdorferi s.l. selectively expresses outer surface proteins (Osps) in the midgut of the unfed tick that are not expressed in the vertebrate host and vice versa (Fikrig et al., 1992; Schwan et al., 1995; de Silva et al., 1999; Fingerle et al., 2000). The most abundant and best studied of the more than 150 (putative) Osps is OspA, followed by OspB, C, D, E, F, pG, VlsE, p66 and a few others (Wallich et al., 1995; Fraser et al., 1997; Zhang et al., 1997; Casjens et al., 2000; Nordstrand et al., 2000). OspA has been shown to act as an adhesin that binds to midgut epithelial cells of ixodid ticks, suggesting an important function of this lipoprotein in tick–spirochaete interactions (Pal et al., 2000). The first step in the transmission of B. burgdorferi s.l. is the attachment of the infected tick to the host. Most ticks of the genus Ixodes feed for 2–7 consecutive days depending on the developmental stage. The uptake of blood triggers B. burgdorferi s.l. to multiply in the midgut of the tick (Piesman et al., 1990). A fraction of the bacterial population present in the tick penetrates the midgut barrier and invades the haemocoel. This process requires the use of the hostderived plasminogen activation system, a mechanism whereby the spirochaetes acquire extracellular proteolytic activity from the host (Fuchs et al., 1994, 1996; Coleman et al., 1995, 1997; Coleman and Benach, 1999). A number of spirochaetal polypeptides, including OspA, have been identified as receptors of plasminogen. During the blood meal of the tick, B. burgdorferi s.l. starts to undergo antigenic phase variation at a variety of loci. In particular, an increasing fraction of the spirochaete population in the tick expresses OspC (Schwan et

B. burgdorferi in the Vertebrate Host

119

al., 1995; de Silva and Fikrig, 1997; de Silva et al., 1999; Fingerle et al., 2000). It has been proposed that only the subpopulation of spirochaetes that express OspC escapes the midgut lumen. This proposition is supported by findings that antibodies to OspC block the migration of spirochaetes from the midgut to the haemocoel (Gilmore and Piesman, 2000). From the salivary glands, B. burgdorferi s.l. is delivered into the feeding lesion that the tick has created.

Dissemination in the host Once inoculated into a host, B. burgdorferi s.l. must disseminate in order to survive in the long term (basic reproduction number, R0 > 1) (Anderson and May, 1991). As a first step in dissemination through the vertebrate host, B. burgdorferi s.l. adheres to host tissue by a variety of binding mechanisms. Most notably, spirochaetes bind to platelets via integrin αIIbβ3 (Coburn et al., 1998) and to endothelial cells via integrins αvβ1and α5β1 (Boeggemeyer et al., 1994; Schaible et al., 1994; Sellati et al., 1996). B. burgdorferi s.l. also adheres to glycosaminoglycans. The recognition of glycosaminoglycans seems to vary amongst different B. burgdorferi s.l. strains, perhaps explaining cell-type specific binding and tissue tropism (Parveen et al., 1999). These in vitro observations are consistent with in vivo experimental, clinical and population genetic studies on B. burgdorferi s.l. that have provided evidence that the bacteria are differentially invasive (Seinost et al., 1999). Invasive strains disseminate from the feeding lesions and infiltrate other host tissues, whereas non-invasive strains remain localized to the feeding lesions and may never infect the host systemically. Dissemination from the site of inoculation into adjacent sites of the dermis is a relatively slow process, even in vertebrate hosts that are highly reservoir-competent for a given B. burgdorferi s.l. strain. For example, rodents require a period of 2–4 weeks to develop a systemic infection with B. burgdorferi sensu stricto (s.s.) (Piesman, 1988; Shih et al., 1992). The kinetics of dissemination suggests that haematogenous spread of B. burgdorferi s.l. is less important than extravascular migration. This is further corroborated by the finding that B. burgdorferi s.l. binds to decorin, a collagen-associated extracellular matrix proteoglycan of the host tissue, mediated by two adhesins of B. burgdorferi s.l., the decorin-binding proteins, DbpA and DbpB, which are expressed within the vertebrate host (Guo et al., 1995). Furthermore, B. burgdorferi s.l. interacts with fibroblasts and fibronectin as well as with fibrocytes in vitro (Klempner et al., 1993; Grab et al., 1998, 1999). The interaction with fibrocytes may be a mechanism of immune evasion, thereby facilitating persistence in the infected host. B. burgdorferi s.l. has been shown to be phagocytosed by professional phagocytes via a mechanism termed ‘coiling phagocytosis’ (Rittig et al., 1998). It is, however, possible that macrophages are involved in dissemination, because spirochaetes occasionally survive in macrophages (Montgomery and Malavista, 1996). Although it can be assumed that invasiveness is governed by a large and as yet unknown variety of molecular interactions, resistance to the complement

120

K. Kurtenbach et al.

system is a major prerequisite for dissemination. The complement system of vertebrates is one of the major effector components of defence against microbes (Sim and Dodds, 1997). As part of the innate immune system it can rapidly respond to microorganisms, long before specific antibodies are generated. The complement system is composed of more than 30 proteins, which interact with each other when the system is activated. Microorganisms can activate complement through the classical or the alternative pathway. Both pathways result in the activation of C3 and the binding of C3b to activator surfaces, followed by an amplification loop and formation of the membrane attack complex (C5b–C9). If not repaired, this eventually leads to cell damage and lysis. Discrimination between ‘self’ and ‘non-self’ is achieved by preferential binding of regulatory proteins that control the amplification loop at different stages. On ‘self-surfaces’ factor H has a higher affinity to C3b than factor B. C3b that binds factor H is catabolized by factor I into iC3b. Degradation of C3b prevents the generation of the alternative pathway C3 convertase, C3bBb, which drives the amplification loop (Sim and Dodds, 1997). Microbial pathogens that are invasive have developed mechanisms of resistance to complement in their evolution. A variety of different mechanisms to escape the complement-mediated attack have been described (Lachmann, 1986; Sim and Dodds, 1997). Recently, it has been discovered that complementresistant B. burgdorferi s.l. strains bind the two host-derived complement control proteins: factor H-like protein 1/reconectin (FHL-1) and factor H (Kraiczy et al., 2001a,b). Five proteins of B. burgdorferi s.l. have so far been found to interact with complement control proteins, collectively termed CRASPs (complement regulatory acquiring surface proteins) (Kraiczy et al., 2001b). Amongst the CRASPs, the outer surface lipoprotein OspE has been identified as a specific ligand for factor H, with the binding site localized to the C-terminal short consensus repeat domains 15–20 of factor H, leaving the N-terminal regulatory domains free to carry out their inhibitory activities (Hellwage et al., 2001). At the time of writing, no sequence data for the other CRASPs are available. The best-studied Osp of B. burgdorferi s.l., OspA, does not interact with complement (Hellwage et al., 2001). Binding of host-derived complement control proteins by B. burgdorferi s.l. is an example of exploitation of host molecules for the deflection of immunity (Finlay and Cossart, 1997). The exploitation of host-derived complement inhibitors is thus an efficient mechanism to avoid lysis and, possibly, complement-mediated enhancement of adaptive immunity. As already mentioned, B. burgdorferi s.l. binds to the plasminogen system of the host. The plasminogen system plays a significant role in fibrinolysis (Fuchs et al., 1994; Coleman et al., 1995; Coleman and Benach, 1999). In the presence of specific activators, plasminogen that is bound to bacterial surfaces is readily converted into plasmin (Fuchs et al., 1996). Plasmin is an aggressive proteolytic enzyme with a broad substrate specificity, e.g. it degrades fibrin, fibronectin and, thus, extracellular matrices. In vitro, such bound plasmin substantially enhances the migration of B. burgdorferi s.s. through endothelial monolayers (Coleman et

B. burgdorferi in the Vertebrate Host

121

al., 1995). The plasmin(ogen) system enhances the dissemination of B. burgdorferi s.l. through both skin and organ invasion, allowing the bacteria to avoid fibrinbased immobilization (Coleman et al., 1997). Plasmin is also effectively used by other invasive bacteria and by tumour cells to disseminate in tissues (Coleman and Benach, 1999). It remains to be elucidated whether binding to plasmin(ogen) is variable depending on the combination of B. burgdorferi s.l. genotype and vertebrate species.

Persistence in the host Once a B. burgdorferi s.l. strain has disseminated from the site of inoculation, infections tend to persist in most animals and in humans lifelong if untreated (Gern et al., 1994; Kurtenbach et al., 1994; Zhong et al., 1997). The lack of self-limitation of B. burgdorferi s.l. infections is remarkable as B. burgdorferi s.l. induces strong cellular and humoral immune responses that target a variety of Osps expressed in the host (Roehrig et al., 1992; Golde et al., 1993; Schaible et al., 1993). Both the quality and the quantity of antibodies produced in response to B. burgdorferi s.l. are dependent on the mode of infection. A large number of studies have been conducted that have analysed this phenomenon using laboratory animals as well as natural vertebrate hosts (Roehrig et al., 1992; Gern et al., 1993; Kurtenbach et al., 1994). It was found that infections with B. burgdorferi s.l. induced by tick bites are normally characterized by a lack of antibodies to OspA and B. However, strong and consistent humoral responses, independent of the genetic background of the vertebrate host, are mounted to other lipoproteins such as OspC, DbpA and DbpB, flagellin, BmpA (formerly known as p39) and VlsE. In contrast, intradermal syringe-induced infections using high numbers of cultured spirochaetes (> 104) generated antibodies to OspA and B in the hosts (Schaible et al., 1993; Kurtenbach et al., 1994). Inoculation with high numbers of killed spirochaetes was found to induce a multivalent antibody response similar to that induced by viable spirochaetes, except for antibodies to BmpA. The presence of antibodies to BmpA, therefore, distinguishes between active and inactive infections in the vertebrate host (Golde et al., 1993). Intradermal infections with low numbers of cultured viable spirochaetes, however, induce antibody profiles with similar specificities to those seen after tickinduced infections. For this reason it has been suggested that the quality of the antibody response depends on the dose of antigen injected (Gern et al., 1993; Schaible et al., 1993). Subsequent studies revealed the mechanism underlying this observation: antigenic phase variation (‘switch on–off’) of B. burgdorferi s.l. Thus, the organisms delivered by the tick are antigenically distinct from cultured forms (Schwan et al., 1995; de Silva and Fikrig, 1997; de Silva et al., 1999; Ohnishi et al., 2001). The stages of B. burgdorferi s.l. delivered by the tick appear to be preadapted to the host environment and do not, or only at low abundance, express OspA and B. OspA and B are thus ‘concealed’ antigens. Consequently, these lipoproteins are only occasionally, if at all, recognized by the immune system of

122

K. Kurtenbach et al.

naturally infected hosts. The finding that midgut-specific stages of B. burgdorferi s.l. selectively express OspA provides the rationale behind the recently approved transmission-blocking vaccine (LYMErix, SmithKline Beecham) that targets OspA in the feeding tick. Antibodies taken up by the feeding tick kill the pathogens before they are delivered into the hosts through the salivary glands (Fikrig et al., 1992). Irrespective of the mode of infection and the antibody profile, B. burgdorferi s.l. infections, once successfully introduced, persist in the host. Several concepts and hypotheses have been proposed to explain the persistence of B. burgdorferi s.l. in the presence of specific immunity. One hypothesis is that B. burgdorferi s.l. resides in immunoprivileged sites of the host, thereby escaping effective clearance by the immune system (Nordstrand et al., 2000). Recently, an alternative working model of transmission has been proposed, suggesting that B. burgdorferi s.l. populations avoid sterile immunity via a novel pattern of antigenic phase variation (Ohnishi et al., 2001). According to this model, the tick delivers an antigenically heterogeneous population of B. burgdorferi s.l. into the host of which only the subpopulation that is depleted of OspA and OspC disseminates. Thus, OspCspecific immune responses as seen in natural infections would be a by-product that do not target the infectious forms of B. burgdorferi s.l. Alternatively, naturally induced OspC-specific antibodies may target B. burgdorferi s.l. in situ, but may be unable to lyse the bacteria in the host (Zhong et al., 1997, 1999). Phylogenetic and experimental studies suggest that ospC and other genes that encode in vivo-expressed antigens recombine genetically (Jauris-Heipke et al., 1995; Zhang et al., 1997; Stevenson et al., 1998; Wang et al., 1999). For example, the gene-tree topologies of ospC are disconcordant with the gene trees based on the ospA or the fla gene (Figs 5.1–5.3), suggesting horizontal gene transfer between different strains of B. burgdorferi s.l. at the ospC locus. Genetic recombination is supported by a major study analysing the genetic (in-) stability of the ospC gene in experimental chains of tick–host infections using Ixodes scapularis and mice (Ryan et al., 1998). In contrast, in other studies analysing genetic stability of needle infections in mice, recombination was not observed (Hodzic et al., 2000). It is, therefore, possible that genetic recombination at the ospC locus requires the natural transmission cycle involving the invertebrate vector. For the vlsE locus, recombination has been observed both in the feeding tick and in the infected host (Zhang et al., 1997). Various genetic mechanisms have been proposed for the vlsE system, in particular intragenic recombination and complex DNA rearrangements similar to those seen in relapsing fever spirochaetes (Zhang et al., 1997; Casjens et al., 2000). Altogether, genetic recombination of B. burgdorferi s.l. genes is an attractive explanation for persistence in the host, as recombination may lead to the emergence of new genotypes that evade effective immunity. In relapsing fever spirochaetes such as Borrelia hermsii, recombination generates new variable membrane protein (vmp) genotypes in the course of infection. New genotypes that are not yet recognized by the host are positively selected, which leads to a temporarily high spirochaetaemia that is later cleared by strong

B. burgdorferi in the Vertebrate Host

123

B. garinii 2, 78

ospA

78

B. garinii ZQ1* B. garinii A87s, VSBM B. garinii Ip1201 B. garinii A77c B. garinii PBi* B. garinii A01c

98

B. garinii A19s B. garinii A91c B. garinii A91s 91

B. garinii 3, 4, 7, 28 B. garinii 31SB, 32SB B. garinii VSDA B. garinii 30SB B. garinii JEM4* B. garinii 11

B. garinii 5

98

ox

92

B. garinii 9, ZQ1 B. garinii Mafra t4, Mafra t20

88

B. garinii Ip89, C8 B. japonica H014*

100

B. japonica H014 B. valaisiana Mafra t6 100

B. burgdorferi s.s. ZS7,* B. burgdorferi s.s. B31 B. bissettii DN127* B. afzelii VS461*, ACA1*

100

B. afzelii ACA1 100

B. turdi Ya501* B. valaisiana Am501* B. valaisiana A14S* 100 100

B. lusitaniae PotiB3* B. lusitaniae PotiB1

B. lusitaniae Mafra PF3 100

B .andersonii 19857* B. andersonii 19857 B. valaisiana 6, 51

100

B .valaisiana VS116*, M19*, UK* B. valaisiana UK

0.05 substitutions/site

Fig. 5.1. Midpoint rooted maximum likelihood tree based on ospA sequences derived from Ixodes uriae ticks (shown in bold), various Borrelia burgdorferi s.l. isolates, and sequences downloaded from GenBank (indicated with an asterisk *). Bootstrap values under 70 are not shown. The bar indicates nucleotide substitutions per site.

124

K. Kurtenbach et al.

B. garinii 2, 78

ospA

78

B. garinii ZQ1* OspA serotype 5

B. garinii A87s, VSBM B. garinii Ip1201 B. garinii A77c

20047 ribotype OspA serotype 6

B. garinii PBi* B. garinii A01c 98

OspA serotype 4 resistant to rodent

B. garinii A19s B. garinii A91c

and human complement

B. garinii A91s 91

B. garinii 3, 4, 7, 28 B. garinii 31SB, 32SB B. garinii VSDA B. garinii 30SB B. garinii JEM4* B. garinii 11

B. garinii 5

98 92 88

OspA serotype 3 resistant to avian complement

OX

B. garinii 9, ZQ1 B. garinii Mafra t4, Mafra t20 B. garinii Ip89, C8

NT 29 ribotype

0.05 substitutions/site

Fig. 5.2. As Fig. 5.1 but showing only B. garinii sequences and their correlation with OspA serotypes and ribotypes.

and specific humoral immunity, followed by the rise of another new variant. Such an extreme form of frequency-dependent selection mediated by strong adaptive immunity does not seem to operate in B. burgdorferi s.l. infections (Qui et al., 1997). Here, the spirochaetal load in the vertebrate host seems to be very low at any time of infection, as indicated by results based on PCR, culturing and histology. The low number of B. burgdorferi s.l. in the host indicates that, despite persistence and the failure to clear the infection, spirochaetemia is under some, probably multifactorial, control involving innate and adaptive components of immunity. This is corroborated by a recent study using a migratory bird species as model, showing that B. burgdorferi s.l. infections can be reactivated in birds during experimentally induced migratory restlessness, whereas spirochaetes could not be recovered by culturing from sedentary birds (Gylfe et al., 2000). Upon experimental induction of physiological restlessness, spirochaetes could be reisolated successfully from the birds. The findings suggest that control mechanisms in the host may break down under certain conditions. At present, the mechanisms of reactivation of B. burgdorferi s.l. infection in the migratory bird system are unclear.

B. burgdorferi in the Vertebrate Host

ospC

100

125

B. afzelii ACA1, * B. burgdorferi s.l. B31*

74

100

71

B. burgdorferi s.s. ZS7, * B. bissettii DN127*

B. garinii PBi* 81

B. garinii A01c OspA serotype 4

B. garinii A19s 100

B. garinii A91c &s, A94s, A76s, A104s B. garinii Ip1201 B. garinii Mafra t4, Mafra t20, VSBP

100

B. garinii Ip560 B. garinii NT29

100

B. japonica HO14, * B. valaisiana A14s* B. afzelii B. tanukii 100

B. valaisiana M19*, VS116*

B. garinii VS102, 20047* 78

100

B. garinii Far02* B. garinii Far03* B. garinii A77s

97 80

B. garinii A77c B. garinii Ip90* B. garinii Ip89, C8 B. garinii ZQ1 B. garinii Mafra t86

92

100 70

B. valaisiana UK, * B. lusitaniae Mafra PF3

B. garinii A87s B. garinii VSDA

0.05 substitutions/site

Fig. 5.3. Midpoint rooted maximum likelihood tree based on ospC sequences derived from a pelagic habitat (shown in bold), various terrestrial Borrelia burgdorferi s.l. isolates, and sequences downloaded from GenBank (indicated with an asterisk). Bootstrap values below 70 are not shown. The bar indicates nucleotide substitutions per site. This gene tree is disconcordant with the ospA tree, indicating genetic recombination. The Mafra strains t4 and t20, which cluster together with OspA serotype 3 strains at the ospA locus, display close genetic similarity with hyperinvasive OspA serotype 4 strains at the ospC locus. As the ospC locus is a marker for invasiveness, the finding suggests that these bird-associated strains from Portugal (as indicated by the ospA sequences) resemble OspA serotype 4 strains in terms of pathology.

126

K. Kurtenbach et al.

The number of B. burgdorferi s.l. cells residing in the host varies with the organ or tissue, suggesting a limited degree of tissue tropism that differs amongst the genotypes of B. burgdorferi s.l. (van Dam et al., 1993; Pennington et al., 1997). For example, in humans Borrelia afzelii seems to be associated with skin, Borrelia garinii with nervous tissue and B. burgdorferi s.s. with joints. As the genome project of B. burgdorferi s.l. has found evidence for the presence of chemotaxis genes, the mechanisms underlying differential tissue tropism may, apart from differential binding patterns to host cells, also be related to allelic variation of chemotaxis genes (Casjens et al., 2000). In conclusion, while innate and adaptive components of natural immunity fail to clear B. burgdorferi s.l. from the host, host immunity substantially suppresses spirochaetemia.

Transmission from the host to the tick As in transmission from the vector to the host, acquisition of spirochaetes by a tick occurs during the blood meal. Ticks may acquire spirochaetes from a variety of vertebrate host species. Murine hosts are the best studied, and a wealth of information indicates that mice remain infectious lifelong, irrespective of their genetic background (Donahue et al., 1987; Gern et al., 1994; Kurtenbach et al., 1994; Hofmeister et al., 1999a,b; Humair et al., 1999). In contrast to tick-to-host transmission, which requires time, B. burgdorferi s.l. is normally more rapidly acquired by a tick from an infected host (Piesman, 1991). However, in order for B. burgdorferi s.l. to be transmitted transstadially, the blood meal must be completed, because below a certain threshold partially fed ticks do not moult (Dizij and Kurtenbach, 1995). Not all ticks acquire a spirochaete infection from an infected host during any one blood meal. Depending on the system studied, the probability of spirochaete acquisition is highly variable (Donahue et al., 1987; Piesman, 1988, 1993b; Gern et al., 1994; Kurtenbach et al., 1994, 1995; Humair et al., 1999). Transmissibility from the host to the tick is unlikely to be simply correlated with spirochaete burden, as the numbers of spirochaetes residing in the skin of the vertebrate host always seem to be low. An increasing body of evidence indicates that transmissibility of B. burgdorferi s.l. from the host to the tick is determined by innate and, to some degree, naturally acquired immunity (Kurtenbach et al., 1995, 1998c; Qui et al., 1997; Hu et al., 2001). In addition, for a variety of animal models there is evidence that transmission is influenced by immunological processes induced by the tick (Dizij and Kurtenbach, 1995). Experimental studies, furthermore, show that the efficiency with which B. burgdorferi s.l. is transmitted from infected hosts to feeding ticks is dependent on the mode of spirochaete inoculation (Gern et al., 1994; Kurtenbach et al., 1994). For example, studies in a variety of natural and laboratory mouse models reported much higher parameter values of host-to-tick transmission coefficients after natural (i.e. tick-borne) infections than after infections induced by needle

B. burgdorferi in the Vertebrate Host

127

inoculation with high numbers of cultured spirochaetes. It has been demonstrated that transmission is inversely correlated with the amount of OspA-specific antibodies defined by the monoclonal antibody LA-2 (Fikrig et al., 1992; Kurtenbach et al., 1994, 1997). Once a tick has completed its blood meal, B. burgdorferi s.l. is, in permissive systems, retained throughout the moult and transmitted to the next developmental stage. This stage, usually the nymph, can then act as a vector, thus completing the B. burgdorferi s.l. transmission cycle. Unlike the kinetics of protein expression and migration patterns of B. burgdorferi s.l. in the feeding vector tick, antigenic phase variation and possible migration during acquisition of spirochaetes from hosts by ticks are not well studied.

Pathogenesis of Lyme borreliosis LB primarily affects humans and some domestic animals, whereas natural reservoir hosts do not appear to develop disease. However, it is possible that some natural hosts do develop mild disease manifestations that readily escape detection. It is difficult to assess disease in natural hosts as they are normally outbred and therefore polymorphic. Very few natural host species have so far been analysed for pathogenesis in experimental studies, and in the field it is even more difficult to detect disease caused by B. burgdorferi s.l. The main reason for this difficulty is that natural hosts are normally carriers of a variety of other endoparasites (such as other bacteria, protozoa or helminths) and ectoparasites (such as lice, fleas and ticks) that may confound rare and/or mild symptoms of disease due to B. burgdorferi s.l. The genome of one strain of B. burgdorferi s.l. has recently been sequenced fully, now also including most of the linear and circular plasmids (Fraser et al., 1997; Casjens et al., 2000). It was hoped that the molecular basis of pathogenicity of B. burgdorferi s.l. would be identified in the course of this genome project. Most interestingly, the project failed to identify any virulence factors or pathogenicity islands in B. burgdorferi s.l. Furthermore, apart from the lack of any defined virulence factors, more than 40% of the open reading frames located on the chromosome and more than 80% of the open reading frames located on the plasmids did not show any similarity with sequences found anywhere else in nature. Therefore, putative proteins inferred from the genome cannot yet be assigned to any biological function. The unique genome and obscure biology render B. burgdorferi s.l. one of the least understood and most interesting pathogens of humans and animals. Why do humans and some domestic and laboratory animals develop LB? The answer to this question can only come from clinical studies, from experimental studies using defined animal models and from functional genomics. Many laboratory animal disease models have already been analysed in detail since the 1980s. The advantage of such models in pathogenesis research is that they are genetically and microbiologically defined. This allows more precise

128

K. Kurtenbach et al.

distinction between host-derived factors and spirochaete-associated factors in the onset of disease. The most important laboratory animal species and strains that have been investigated include hamsters (Schmitz et al., 1989), gerbils (Preac-Mursic et al., 1990), guinea pigs (Sonnensyn et al., 1993), laboratory mice (Schaible et al., 1989; Barthold et al., 1990; Bockenstedt et al., 1997), rats (Barthold, 1995), rabbits (Barthold et al., 1988), dogs (Appel et al., 1993) and monkeys (Philipp et al., 1993). Amongst these animal models, inbred and recombinant strains of mice have been particularly valuable in providing insight into the pathogenesis of spirochaete infection. Therefore, in this section emphasis will be placed on laboratory mouse models of LB. Some strains of mice, like humans, develop multisystemic disease, mainly arthritis and cardiovascular symptoms (Schaible et al., 1989; Museteanu et al., 1991). One of the major findings using mouse models was that disease seems to be mediated by spirochaete-induced inflammatory responses that are often out of proportion to the spirochaetal load. This suggests that B. burgdorferi s.l. is able to trigger amplifying inflammatory processes that cause the disease manifestations (Nordstrand et al., 2000). In vitro studies showed that B. burgdorferi s.l. is able to activate macrophages and to induce oxygen radicals and NO (Modelell et al., 1994) as well as a variety of proinflammatory cytokines like interleukin (IL)-1β, IL-6 and tumour necrosis factor-α (Ma and Weiss, 1993; Radolf et al., 1995). Furthermore, the induction of a large variety of chemokines, such as IL-8 (Ebnet et al., 1996a,b; Burns et al., 1997), which activate various types of white blood cell, suggests a crucial role in pathogenesis (Sprenger et al., 1997). Although no virulence factors could be inferred from the genome itself, it is likely that the lipoproteins and glycoproteins associated with the outer surface of B. burgdorferi s.l. (many of which are encoded by genes located on plasmids) trigger the disease-causing processes (Radolf et al., 1995). For example, preparations of these lipoproteins directly stimulate macrophages and B cells and provide costimulatory signals for T cells (Simon et al., 1995; Knigge et al., 1996). This suggests that pathological processes are driven by interactions between the outer surface lipo(glyco)proteins of the spirochaetes and infiltrating and tissue-specific cells (Simon et al., 1993; Nordstrand et al., 2000). In vivo studies using immunocompetent (Barthold et al., 1990) and immunocompromised strains of mice (Schaible et al., 1989) have provided evidence that pathogenesis of B. burgdorferi s.l. is linked to the genetic background of the animal model and to the genotype of B. burgdorferi s.l. (Pennington et al., 1997). These studies have also established that the adaptive immune response is a critical element in pathogenesis, being either ameliorative or pathogenic. A variety of disease-susceptible and resistant mouse models have been developed and investigated. Outbred and inbred mice are all susceptible to B. burgdorferi s.l. infection. However, susceptibility to disease differs between inbred strains of mice representing different haplotypes of the major histocompatibility complex (MHC). This reinforces the notion that infection and disease have to be dissociated in LB (Thomas et al., 2001). The role that acquired immunity plays in

B. burgdorferi in the Vertebrate Host

129

pathogenesis is underlined by the fact that the various H-2 haplotypes of mice differ in their cellular immune response to B. burgdorferi s.l. For example, the immunocompetent mouse strains BALB/cByJ (H-2d) and C57BL/6J (H-2b) have been found to be relatively disease-resistant to B. burgdorferi s.s., whereas C3H/HeJ (H-2k) or SWR/J (H-2q) mice were found to develop severe disease symptoms after infection with the same pathogenic borrelial strains (Simon et al., 1993; Barthold, 1995). For this reason the C3H mouse strain has become one of the most widely used disease-susceptible mouse strains in LB pathogenesis research (Oteo et al., 1998). Of all the MHC haplotypes analysed, H-2d strains of inbred mice seem to be the most disease-resistant. The conclusion that acquired immunity controls disease is further supported by the finding that some of the disease-resistant MHC haplotypes, in particular H-2d strains such as BALB/C, are susceptible to disease when infected at a very young age, i.e. at an age at which they are still immunocompromised (Barthold et al., 1990). On the other hand, studies using mice with severe combined immunodeficiency (scid) syndromes have shown that disease is also controlled by genes that regulate innate susceptibility. Scid mice lack any functional B or T cells and are therefore not able to produce immunoglobulins. For example, comparison of the immunocompromised strains C3H.scid, C.B.-17.scid, RAG-1D/D or RAG-2 D/D showed that the first two strains do develop severe arthritis, in contrast with the RAG strains, which do not (Schaible et al., 1989; Simon et al., 1993). The relative importance of acquired versus innate immunity to control or mediate disease remains a matter of debate. Apart from the genetic background of the host, there seem to be intrinsic differences amongst the different genospecies of B. burgdorferi s.l. in disease induction. In humans, for example, of the 11 genospecies delineated so far, only B. afzelii, B. garinii and B. burgdorferi s.s. have unambiguously been associated with disease (van Dam et al., 1993). However, some genospecies previously considered to be apathogenic, such as Borrelia lusitaniae, have recently been found to cause disease in C3H mice (N. Zeidner, Fort Collins, USA, personal communication). On the other hand, B. burgdorferi s.l. isolates originally pathogenic may lose entire plasmids after multiple in vitro passages and this has been associated with the loss of pathogenic potential of the bacteria (Schwan et al., 1988; Siebers et al., 1999; Casjens et al., 2000). Pathogenesis research in LB has passed into the genomic era and it is likely that functional genomics will shed more light on disease-inducing molecules of B. burgdorferi s.l. in the near future (Fraser et al., 1997; Nordstrand et al., 2000).

Diversity of Tick–Host–Spirochaete Interactions The processes governing transmission, persistence and pathogenesis of B. burgdorferi s.l. are surprisingly diverse. We are only beginning to understand the extent and impact of this diversity. First of all, the bacterium displays considerable genetic and phenotypic diversity. Secondly, the range of vertebrate

130

K. Kurtenbach et al.

species found to be susceptible to infection and also transmission-competent for one or more strains of B. burgdorferi s.l. is increasing. Of equal interest are those vertebrate species that are not transmission-competent because they may reduce the basic reproduction number (R0 ) of B. burgdorferi s.l. Thirdly, the tick–host relationships differ substantially from each other depending on the species involved, with the acquisition of resistance to ticks on one side of the biological range and enhanced tolerance to repeated tick bites on the other (Dizij and Kurtenbach, 1995). These diverse interactions determine the life cycle and ecology of B. burgdorferi s.l.

Diversity of B. burgdorferi s.l. Since 1992, B. burgdorferi s.l. is no longer considered to be a single bacterial species, but rather a diverse species complex now comprising 11 named genospecies: B. afzelii, B. andersonii, B. bissettii, B. burgdorferi s.s., B. garinii, B. japonica, B. lusitaniae, B. sinica, B. tanukii, B. turdi and B. valaisiana. Some of the genospecies of B. burgdorferi s.l. display considerable intraspecific diversity. B. garinii appears to be the most polymorphic genospecies (Marti-Ras et al., 1997; Masuzawa et al., 1997; Postic et al., 1998). The phylogenetic and evolutionary relationships amongst the different strains of B. burgdorferi s.l. are not yet fully resolved (Marti-Ras et al., 1997). The degree of clonality is one of the controversial issues discussed (Dykhuizen et al., 1993; Wang et al., 1999; Dykhuizen and Baranton, 2001). Horizontal gene transfer between different spirochaetal strains, and other mechanisms of recombination, may disrupt the clonal frame of B. burgdorferi s.l. (Stevenson et al., 1998; Sung et al., 1998; Wang et al., 1999). However, the rates of genetic recombination of B. burgdorferi s.l. loci are not known. The increasingly recognized genetic and phenotypic diversity of the bacteria has an important impact on the diversity of tick–host–spirochaete interactions discussed in this chapter, and provides a framework for eco-epidemiological studies.

Diversity of vertebrate hosts Globally, more than 100 natural vertebrate host species have been identified as playing a role in the ecology of B. burgdorferi s.l. (Anderson, 1991; Oliver, 1996; Gern et al., 1998). As the physiology, immunology and behavioural ecology of these vertebrates are highly diverse, each animal species (or subpopulation thereof) may contribute differently to the biology of LB. The most extensively studied natural hosts are small mammals of the genera Neotoma and Peromyscus in North America (Mather et al., 1989; Brunet et al., 1995; Oliver, 1996) and Apodemus and Clethrionomys in Eurasia (Kurtenbach et al., 1994, 1995; Humair et al., 1999). Rodents substantially contribute to the basic

B. burgdorferi in the Vertebrate Host

131

reproduction number (R0) of B. burgdorferi s.l. in nature by virtue of their high capacity to infect larval ticks. Perhaps for this reason, rodents have been considered to be the principal reservoir hosts of B. burgdorferi s.l. In recent years, however, it has become apparent that some bird species are also reservoir hosts for B. burgdorferi s.l. (Olsen et al., 1996; Piesman et al., 1996; Humair et al., 1998; Kurtenbach et al., 1998a; Rand et al., 1998; Richter et al., 2000). Furthermore, avian hosts, particularly migratory birds, contribute to the continental and even transhemispheric dispersal of B. burgdorferi s.l. (Olsen et al., 1993, 1995a,b). Many natural vertebrate hosts are unable to infect ticks with B. burgdorferi s.l. and are therefore considered to be reservoir-incompetent. The most prominent example is deer (apparently most, if not all, species) (Telford et al., 1988; Matuschka et al., 1993; Gern et al., 1998; Nelson et al., 2000). For this reason, it has been concluded that deer play a zooprophylactic role in the ecology of LB (Matuschka et al., 1993). However, reservoir incompetence does not necessarily imply zooprophylaxis. Deer often feed large numbers of immature and adult ticks and boost the tick population. As the population density of ticks is an important variable in the equation of the basic reproduction number R0 (Randolph, 1998), mathematical modelling is required to determine whether deer reduce or increase R0 of LB spirochaetes. Host specificity of B. burgdorferi s.l. One of the most important findings in LB research in recent years is that the different genospecies or strains of B. burgdorferi s.l. are maintained in nature by different spectra of vertebrate host species. For example, the woodmouse, Apodemus sylvaticus, is frequently infected with B. afzelii in central parts of Europe (Humair et al., 1995, 1999; Kurtenbach et al., 1998b, 2001). In central parts of North America, woodrats (Neotoma spp.) are infected predominantly with B. bissettii (Maupin et al., 1994; Schneider et al., 2000). In New England, USA, mice of the genus Peromyscus are mainly infected with B. burgdorferi s.s. (Donahue et al., 1987; Mather et al., 1989). In Japan, rodents are predominantly carriers of B. japonica (Nakao et al., 1994a,b; Masuzawa et al., 1997). Seabirds analysed on islands in the North Sea are infected with B. garinii, but not with B. afzelii (Olsen et al., 1993, 1995a). It is evident that a variety of extrinsic ecological factors (such as climate, vegetation or geology) determine the relative role a vertebrate host population plays in the maintenance and geographical distribution of a particular strain of B. burgdorferi s.l. (Kurtenbach et al., 1995; Randolph and Craine, 1995; Kirstein et al., 1997; Randolph, 1998; de Michelis et al., 2000). It has, however, become apparent that intrinsic factors also play important roles in the biology of B. burgdorferi s.l. A growing body of experimental and ecological evidence indicates that different genotypes of B. burgdorferi s.l. differ in their transmissibility between hosts and ticks. This suggests a strong element of host-specificity that is independent of extrinsic ecological factors (Kurtenbach et al., 1998b; Gylfe et al., 1999; Humair

132

K. Kurtenbach et al.

et al., 1999; Richter et al., 2000; Hu et al., 2001). Thus, reservoir competence is variable depending on the combination of spirochaetal strain and vertebrate host species. The overall pattern that has emerged is that small mammals are reservoir-competent for a different set of B. burgdorferi s.l. strains from that of avian hosts. This specific transmission pattern suggests that many genotypes or strains of B. burgdorferi s.l. are adapted to either mammalian or avian hosts but not to both. In contrast, B. burgdorferi s.s. appears to be less specialized and often seems to be infectious for both avian and mammalian hosts. Based on the findings on differential transmission patterns, B. burgdorferi s.l. strains can be divided into three groups: (i) strains that are adapted to small mammals; (ii) strains that are adapted to birds; and (iii) strains that are not specialized (Fig. 5.4). The findings on differential transmissibility and host-specificity of B. burgdorferi s.l. require a cautious use of the terms ‘reservoir competence’, ‘reservoir incompetence’ and ‘zooprophylaxis’ because one vertebrate species may be reservoir-competent for one strain of B. burgdorferi s.l. but reservoir-incompetent for another. The association of spirochaetal genotypes with particular vertebrates has an important implication for the dispersal of B. burgdorferi s.l. Unlike small mammals, which have a migration rate of only 200–300 m per generation, many birds have high rates of migration, a parameter in the population biology and ecology of B. burgdorferi s.l. that has only recently been addressed. High rates of migration of B. garinii genotypes have been found by various research groups (Olsen et al., 1995a; Gylfe et al., 1999). For example, the authors of this chapter have isolated strains of B. garinii with almost identical ospA and flagellin sequences

Bird

Small rodent

B. garinii (NT 29 ribotype; B. garinii (20047 ribotypes; OspA serotypes 3, 5, 6, 7)

B. valaisiana

OspA serotype 4)

B. afzelii B. bissettii B. japonica

B. burgdorferi s.s. Fig. 5.4. Schematic diagram of host-specificity of B. burgdorferi s.l. Avian hosts are reservoir-competent for a different set of strains from that of small rodents.

B. burgdorferi in the Vertebrate Host

133

from questing ticks in Portugal and from seabird ticks (Ixodes uriae) collected on islands in the North Sea (unpublished observations; figs 1–3). Altogether, it is now established that B. burgdorferi s.l. is a multi-host bacterial species complex that is ecologically differentiated (Fig. 5.4). Ecological differentiation, however, is not always consistent with genospecies classification, as a genospecies may comprise divergent ‘ecotypes’. Thus, whilst the current genospecies concept has certain benefits, it obviously lacks precise predictive power in ecological studies of LB. Because of independent transmission, the values for R0 of B. burgdorferi s.l. need to be estimated mathematically for each genospecies. It is likely that the sum of the species-specific values for the basic reproduction number is lower than previously calculated for the species complex as a single unit (Randolph and Craine, 1995; Randolph, 1998). The complement system as a determinant of host-association specificity It is now accepted that the complement system is a major determinant of hostspecificity (i.e. niche adaptation) of B. burgdorferi s.l. (Kurtenbach et al., 1998c; Kuo et al., 2000; Nelson et al., 2000). Therefore, this aspect of the biology of B. burgdorferi s.l. deserves special attention. Cultured strains of B. burgdorferi s.l. are highly heterogeneous with respect to complement-mediated lysis. Brade and co-workers first reported on differences in serum resistance amongst strains of B. burgdorferi s.l. (Brade et al., 1992; BreitnerRuddock et al., 1997). The studies on resistance/sensitivity patterns to complement have since been extended to a large number of B. burgdorferi s.l. strains and different vertebrate species including avian, rodent and ruminant hosts (van Dam et al., 1997; Kurtenbach et al., 1998b) (Tables 5.1 and 5.2). A clear pattern of resistance/sensitivity by B. burgdorferi s.l. to complement has emerged that correlates with patterns of traits that are biologically significant, for example OspA serotype or transmissibility. For example, B. afzelii, a genospecies that is not transmitted by avian hosts, is lysed rapidly in vitro by avian complement. Conversely, B. valaisiana and the subtypes of B. garinii that are transmitted by avian species (20047 or OspA serotypes 3, 5, 6, 7) are resistant to avian complement. Invasive strains of B. garinii that belong to the OspA serotype 4 group of isolates from cerebrospinal fluid of patients and those that are related to the Asian ribotype NT29 are relatively resistant to human and rodent complement, but sensitive to avian complement (Kurtenbach et al., 1998c; K. Kurtenbach et al., unpublished observations). Various studies have attempted to identifiy the mechanisms that confer resistance to complement in the Borrelia system. All B. burgdorferi s.l. strains invariably bind C3. It is now established that resistant and sensitive strains differ with respect to deposition of C5b–C9, the membrane attack complex. As already described, resistance to complement is mediated by species-specific binding of two complement control proteins: factor H-like protein 1/reconectin and factor H. Species-restricted binding of complement control proteins is mediated by

134

K. Kurtenbach et al.

Table 5.1. Resistance of Borrelia burgdorferi sensu lato to host complement. B. burgdorferi s.s

B. afzelii

B. japonica

B. lusitaniae

B. valaisiana

B. garinii 20047

NT29

Vole Hamster Squirrel Yellow-necked mouse Woodmouse Pheasant Guillemot Blackbird Lizard (Lacerta) Sheep Horse Pig Cow Deer

Fully resistant toresistant host complement. Fully to host complement. Partially resistant toresistant host complement. Partially to host complement. Not resistant (i.e. sensitive) complement. Not resistant (i.e.tosensitive) to complement.

B. burgdorferi s.l. CRASPs (Hellwage et al., 2001; Kraiczy et al., 2001a,b). Since OspE, a member of a large, plasmid-borne multigene family (see Bergström et al., Chapter 3) is one of the CRASPs, it is possible that the other CRASPs are also members of this family. A model of transmission The studies on resistance/sensitivity patterns were all performed in vitro using cultured strains. In terms of protein expression, cultured forms resemble the stages of B. burgdorferi s.l. found in the midgut of the tick (de Silva and Fikrig, 1997), suggesting that the CRASPs that serve as specific ligands for host-derived complement control proteins are also expressed in the midgut of the tick. Although only data on OspE exist (Hellwage et al., 2001), this suggestion is strongly corroborated by the precise match of complement-resistance patterns found in vitro with patterns of transmission (Kurtenbach et al., 1998b,c). Here a working model is proposed that depicts how complement may critically operate in the transmission cycle of B. burgdorferi s.l. Infected ticks that contain B. burgdorferi s.l. in the midgut take up host complement in large quantities during the blood meal along with other host-derived molecules such as plasminogen. It

B. burgdorferi in the Vertebrate Host

135

Table 5.2. Bacteriolysis of B. burgdorferi sensu lato by different concentrations of untreated sera after 2, 4 and 22 h of incubation.a

Animal species/spirochaete Mouse/B. garinii (ZQ 1)

Pheasant/B. afzelii (ACA 1)

Sheep/B. burgdorferi s.s. (ZS 7)

Deer/B. garinii (ZQ 1)

Serum conc. (%) 10 25 50 10 25 50 10 25 50 10 25 50

% bacteriolysis at incubation time (h) 2 4 22 5 42 >95 0 11 17 5 9 14 72 100 100

24 100 100 11 19 28 13 12 18 78 100 100

61 100 100 31 87 100 21 43 84 100 100 100

a

For each combination of animal species and spirochaetal strain, the per cent mortality of spirochaetes is given for one representative untreated serum sample (from Kurtenbach et al., 1998c).

has been shown, at least for I. ricinus, that complement is active in the feeding tick (Papatheodorou and Brossard, 1987) despite the possible uptake of salivary anti-complement proteins (Valenzuela et al., 2000). As seen in vitro, the model predicts that ingested complement interacts with the midgut stages of B. burgdorferi s.l. Resistant strains, for example B. afzelii encountering rodent complement during the blood meal of a tick on a mouse, specifically bind rodent factor H, conferring resistance to rodent complement. Consequently B. afzelii survives the blood meal, escapes the midgut lumen and invades the salivary glands from where the bacteria are delivered to the host (Fig. 5.5). Complement-resistant strains are also transmitted transstadially through the moult to the next developmental stage of the tick. In contrast, sensitive strains, for example B. afzelii that encounter deer or avian complement, activate the complement cascade, which kills the bacteria in the midgut before they are transmitted to the host (Figs 5.5 and 5.6). In addition, elimination of complement-sensitive spirochaetes from the midgut prevents transstadial transmission of the bacteria to the next developmental stage of the tick. The model proposed here suggests that the spirochaetes are selectively vulnerable to complement in the midgut of the feeding tick. The differential lysis or survival of the bacteria within the midgut renders the tick’s midgut a bottleneck in terms of transmission of B. burgdorferi s.l. Selective elimination/retention of B. burgdorferi s.l. in the midgut of the tick may manifest itself in the replacement of B. burgdorferi s.l. strains during development from the nymph to the adult tick, as each developmental stage may feed on a different host. For example, a

136

K. Kurtenbach et al.

Tick

Host

p C*

p

p H

C*

H

H H

H

C*

H

H

C* p

p

H

H

C*

Fig. 5.5. Schematic figure depicting the proposed model of transmission of B. burgdorferi s.l. through the tick into the host. Plasminogen (P), complement (C*) and complement control proteins (H) are taken up by the feeding tick. In permissive systems, B. burgdorferi s.l. binds complement control proteins in the midgut (broken line) and is protected against complement-mediated lysis. Dissemination through the tick into the salivary glands takes at least 48 h and is facilitated by binding of plasminogen/plasmin. Spirochaetes that have undergone antigenic phase variation (i.e. downregulation of OspA; solid line) are eventually delivered to the host. In non-permissive systems, spirochaetes do not bind complement control proteins, resulting in the formation of the membrane attack complex and lysis of bacteria in the midgut before they are transmitted to the host. In systemically infected ticks, spirochaetes are already present in the salivary glands at the beginning of the blood meal. In such ticks, complement-mediated lysis of spirochaetes in the midgut may not prevent rapid transmission to the host.

questing nymph infected with B. afzelii that feeds on a bird that is infected with B. garinii (e.g. 20047 or OspA serotype 6) is likely to lose the B. afzelii infection but acquire B. garinii. In fact, in a study from Europe mixed infections of B. afzelii and B. garinii in individual questing adult ticks were found significantly less frequently than expected under neutralist assumptions (Kurtenbach et al., 2001). However, the working model presented here needs to be reconciled with a conflicting observation. Wild rodents have occasionally been found to be carriers of 20047-like variants of B. garinii and OspA serotype 6 strains (Kurtenbach et al., 1998b; K. Kurtenbach et al., unpublished observations). As these strains are highly sensitive to rodent complement in vitro, one would expect that vector ticks carrying such variants would lose their infection. This raises the important question: how do hosts whose complement system lyses B. burgdorferi s.l. in vitro become occasionally infected with such strains in the first place? According to

B. burgdorferi in the Vertebrate Host

1)

Tick *

137

Host *

Tick *

Tick *

2)

Tick *

//

Host

//

Tick

Host *

//

Tick

Tick

3)

Tick

syst.

Tick Fig. 5.6. Schematic diagram of possible transmission routes and blockades of B. burgdorferi s.l. from infected vector ticks to hosts and recipient ticks. An asterisk means the tick or host is infected. (1) Permissive system with horizontal and vertical transmission through the cycle. Example: B. afzeliiinfected tick feeding on a rodent. (2) Non-permissive system. No horizontal transmission from the infected tick (midgut infection) to the host and no vertical transmission to the next developmental stage of the tick (i.e. elimination of midgut infection in vector). Example: tick infected with B. afzelii feeding on an avian host or deer. (3) Semi-permissive system. Horizontal transmission to the host by systemically infected vector ticks and possible vertical transmission through the moult, but no further transmission to recipient ticks. Example: ticks systemically infected with B. garinii (e.g. OspA serotypes 5, 6) may infect rodents. However, rodents carrying such strains do not infect feeding ticks.

the model of transmission presented here, B. burgdorferi s.l. should be killed in the midgut of the tick, blocking any further development of the bacteria in the tick, thereby preventing transmission to that host. The answer may come from systemically infected ticks (Fig. 5.6). Systemically infected ticks may carry forms of B. burgdorferi s.l. in the salivary glands that are already pre-adapted to the host environment. We propose that B. burgdorferi s.l. delivered by systemically infected ticks may escape complement-mediated destruction. It remains to be seen whether resistance to complement of B. burgdorferi s.l. in the host is mediated by the CRASPs identified in cultured strains or by other mechanisms. Here it is hypothesized that resistance to complement of spirochaetes within the host is not species- or strain-restricted and that it is conferred by structural protection (e.g. glycosylation) and not by exploitation of host-derived complement control

138

K. Kurtenbach et al.

proteins. Functional genomics will provide further insight into the complex interactions of B. burgdorferi s.l. and the complement system. The discovery that genotypes of B. burgdorferi s.l. are eliminated from the feeding tick by a host-derived factor adds a novel parameter to models of the life cycle of B. burgdorferi s.l. Until recently, most models of transmission assumed that the coefficients of transstadial transmission of B. burgdorferi s.l. (β T-T) were close to 1 (Randolph and Craine, 1995; Randolph, 1998). It was assumed that transstadial transmission is highly and invariably efficient, irrespective of the source of the blood meal or the spirochaetal genotype. In Fig. 5.6, the three possible transmission pathways of B. burgdorferi s.l. are summarized schematically. The observation that sensitive genotypes are eliminated by host complement taken up by the tick during the blood meal is reminiscent of the elimination of spirochaetes by OspA-specific antibodies. The model of transmission presented here does not deny the importance of other factors that may determine transmission of B. burgdorferi s.l. through ticks and hosts. In fact, it is highly probable that a large number of different and as yet unknown specific molecular interactions have evolved that define a permissive condition for transmission. However, specific binding of B. burgdorferi s.l. proteins to host-derived complement control proteins in the tick is a prerequisite of successful transmission. In other words, failure of B. burgdorferi s.l. to bind complement control proteins results in a blockade of transmission. Resistance/ sensitivity patterns of B. burgdorferi s.l to complement can be used as an in vitro tool to predict or rule out reservoir competence of hosts for B. burgdorferi strains whose ecology is unknown, for example European strains of B. bissettii and B. lusitaniae. The prediction that OspA serotype 4 strains of B. garinii are transmissible through rodents to ticks has already been validated (van Dam et al., 1997; Marconi et al., 1999; Hu et al., 2001). In summary, it has now been established that interaction with complement is a critical element in the life cycle of B. burgdorferi s.l. and therefore in the global ecology of LB (Kurtenbach et al., 1998b; Lane and Quistad, 1998; Kuo et al., 2000; Nelson et al., 2000).

Concluding Remarks B. burgdorferi s.l. is a remarkable and unique bacterium. The majority of genes encoding proteins of the outer surface coat do not match anything in the tree of life. The bacterium causes disease, and yet no virulence factors have been found. Very few bacterial cells reside in the vertebrate host, and yet they may trigger pathological reactions that are entirely out of proportion. B. burgdorferi s.l. strongly stimulates the adaptive immune system, and yet it persists in the vertebrate host, often lifelong. It performs a sophisticated form of deflection of the innate immune system for its own benefit. Diversity is created by ‘localized sex’, perhaps in the tick. Its ecological niche is defined by the micro-environment at the interface of tick and host, rather than solely in the host. It is, in evolution-

B. burgdorferi in the Vertebrate Host

139

ary terms, ‘old’ and certainly not an emerging bacterium, and yet was detected only in the 1980s. Adaptive radiation of B. burgdorferi s.l. has enabled this species complex to occupy a large variety of ecological niches. Amongst the genospecies, speciation and niche adaptation are most obvious for B. garinii. This genospecies displays pronounced subspecies structure with clades that are adapted to aviantick transmission cycles and others that are adapted to rodent-tick transmission cycles. It is possible that ospC alleles that mark hyper-invasiveness are horizontally transferred into bird-associated spirochaetal strains. This, along with dramatically changing patterns of avian migration (Berthold, 1998), could pose a threat to public health. Important biological questions remain to be addressed in LB research. Is B. burgdorferi s.l. originally a tick-associated commensal or a microparasite of mammalian, avian or even reptilian species? When in natural history did B. burgdorferi s.l. switch to its present strictly vector-borne lifestyle, relying on the alternating passage between vertebrate host and invertebrate vector? Functional genomics, molecular phylogeny, population genetics coupled with ecology and experimental transmission studies are likely to answer such questions. In some ways, the genome project of B. burgdorferi s.l. has closed doors, but in many ways it has opened new doors. The post-genomic era ahead is likely to become a productive period in LB research.

Acknowledgements We would like to thank Markus Simon, Peter Kraiczy, Volker Brade, Margarida Collares-Pereira, Margarida Santos-Reis, Rainer Wallich, Lise Gern, Patricia Nuttall, Sarah Randolph, Edward Holmes, Olaf Kahl, Nick Ogden, Joseph Piesman, Amy Fagerberg, Michael Donaghy, Milan Labuda, Klara Hanincová and Jeremy Gray for constructive advice. Some of the work presented was supported by the Boehringer Ingelheim Fonds, Stuttgart, the German Academic Exchange Service (DAAD), Bonn, Germany, the Natural Environment Research Council, Swindon, UK, and the Wellcome Trust, London, UK.

References Anderson, J. (1991) Epizootiology of Lyme borreliosis. Scandinavian Journal of Infectious Diseases Suppl. 77, 23–34. Anderson, R.M. and May, R.M. (1991) Infectious Diseases of Humans, Dynamics and Control. Oxford University Press, Oxford, UK, 575 pp. Appel, M.J.G., Allan, S. and Jacobson, R.H. (1993) Experimental Lyme disease

in dogs produces arthritis and persistent infection. Journal of Infectious Diseases 167, 651–664. Barthold, S.W. (1995) Animal models for Lyme disease. Laboratory Investigation 72, 127. Barthold, S.W., Moody, K.D., Trewilliger, G.A., Duray, P.H., Jacoby, R.O. and Steere, A.C. (1988) Experimental Lyme

140

K. Kurtenbach et al.

arthritis in rats infected with Borrelia burgdorferi. Journal of Infectious Diseases 157, 842–846. Barthold, S.W., Beck, D.S., Hansen, G.M., Terwilliger, G.A. and Moody, K.D. (1990) Lyme borreliosis in selected strains and ages of laboratory mice. Journal of Infectious Diseases 162, 133–138. Berthold, P. (1998) Bird migration. Genetic programs with high adaptability. Zoology – Analysis of Complex Systems 101, 235–245. Bockenstedt, L.K., Hidzic, E., Feng, S., Bourrel, K.W., de Silva, A., Montgomery, R.R., Fikrig, E., Radolf, J.D. and Barthold, S.W. (1997) Borrelia burgdorferi strain-specific OspC-mediated immunity in mice. Infection and Immunity 65, 4661–4670. Boeggemeyer, E., Stehle, T., Schaible, U.E., Hahne, M., Vestweber, D. and Simon, M.M. (1994) Borrelia burgdorferi upregulates the adhesion molecules Eselectin, P-selectin, ICAM-1 and VCAM-1 on mouse endothelioma cells in vitro. Cell Adhesion Communication 2, 145–157. Brade, V., Kleber, I. and Acker, G. (1992) Differences of two Borrelia burgdorferi strains in complement activation and serum resistance. Immunobiology 185, 453–465. Breitner-Ruddock, S., Wuerzner, R., Schulze, J. and Brade, V. (1997) Heterogeneity in the complementdependent bacteriolysis within the species of Borrelia burgdorferi. Medical Microbiology and Immunology 185, 253–260. Brunet, L.R., Selitto, C, Spielman, A. and Telford, S.R. III (1995) Antibody response of the mouse reservoir to Borrelia burgdorferi in nature. Infection and Immunity 63, 3030–3036. Burns, M.J., Sellati, T.J., Teng, E.I. and Furie, M.B. (1997) Production of interleukin-8 (IL-8) by cultured endothelial cells in response to Borrelia burgdorferi occurs independently of secretion of IL-

1 and tumor necrosis factor alpha and is required for subsequent transendothelial migration of neutrophils. Infection and Immunity 65, 1217–1222. Casjens, S., Palmer, N., van Vugt, R., Huang, W.M., Stevenson, B., Rosa, P., Lathigra, R., Sutton, G., Peterson, J., Dobson, R.J., Haft, D., Hickey, E., Gwinn, M., White, O. and Fraser, C.M. (2000) A bacterial genome in flux: the twelve linear and nine circular extrachromosomal DNAs in an infectious isolate of the Lyme disease spirochete Borrelia burgdorferi. Molecular Microbiology 35, 490–516. Coburn, J., Magoun, L., Bodary, S.C. and Leong, J.M. (1998) Integrins αvβ3 and α5β1 mediate attachment of Lyme disease spirochaetes to human cells. Infection and Immunity 66, 1946–1952. Coleman, J.L. and Benach, J.L. (1999) Use of the plasminogen activation system by micro-organisms. Journal of Laboratory Clinical Medicine 134, 567–574. Coleman, J.L., Sellati, T.J., Testa, J.E., Kew, R.R., Furie, M.B. and Benach, J.L. (1995) Borrelia burgdorferi binds plasminogen, resulting in enhanced penetration of endothelial monolayers. Infection and Immunity 63, 2478–2484. Coleman, J.L., Gebbia, J.A., Piesman, J., Degen, J.L., Bugge, T.H. and Benach, J.L. (1997) Plasminogen is required for efficient dissemination of B. burgdorferi in ticks and for enhancement of spirochetemia in mice. Cell 89, 1111–1119. van Dam, A.P., Kuiper, H., Vos, K., Widjojokusumo, A., de Jongh, B.M., Spanjaard, L., Ramselar, A.C., Kramer, M.D. and Dankert, J. (1993) Different genospecies of Borrelia burgdorferi are associated with distinct clinical manifestations of Lyme borreliosis. Clinical Infectious Diseases 17, 708–717. van Dam, A.P, Oei, A., Jaspars, R., Fijen, C., Wilske, B., Spanjaard, L. and Dankert, J. (1997) Complement-mediated serum activity among spirochetes

B. burgdorferi in the Vertebrate Host

that cause Lyme disease. Infection and Immunity 65, 1228–1236. Dizij, A. and Kurtenbach, K. (1995) Clethrionomys glareolus, but not Apodemus flavicollis, acquires resistance to Ixodes ricinus (Acari: Ixodidae), the main European vector of Borrelia burgdorferi. Parasite Immunology 17, 177–183. Donahue, J.G., Piesman, J. and Spielman, A. (1987) Reservoir competence of white-footed mice for lyme-disease spirochetes. American Journal of Tropical Medicine and Hygiene 36, 92–96. Dykhuizen, D.E. and Baranton, G. (2001) The implications of low rate of horizontal transfer in borrelia. Trends in Microbiology 9, 344–350. Dykhuizen, D.E., Polin, D.S., Dunn, J.J., Wilske, B., Preac-Mursic, V., Dattwyler, R.J. and Luft, B. (1993) Borrelia burgdorferi is clonal: implications for taxonomy and vaccine development. Proceedings of the National Academy of Sciences, USA 90, 10163–10167. Ebnet, K., Brown, K.D., Siebenlist, U.K., Simon, M.M. and Shaw, S. (1996a) Borrelia burgdorferi activates nuclear factor-kappa B and is a potent inducer of chemokine and adhesion molecule gene expression in endothelial cells and fibroblasts. Journal of Immunology 158, 3285–3292. Ebnet, K., Simon, M.M. and Shaw, S. (1996b) Regulation of chemokine gene expression in human endothelial cells by proinflammatory cytokines and Borrelia burgdorferi. Annals of the New York Academy of Sciences 797, 107–117. Fikrig, E., Telford, S.R. III, Barthold, S.W., Kantor, F.S., Spielman, A. and Flavell, R.A. (1992) Elimination of Borrelia burgdorferi from vector ticks feeding on OspA-immunized mice. Proceedings of the National Academy of Sciences, USA 89, 5418–5421. Fingerle, V., Laux, H., Munderloh, U.G., Schulte-Spechtel, U. and Wilske, B. (2000) Differential expression of outer surface proteins A and C by individual

141

Borrelia burgdorferi in different genospecies. Medical Microbiology and Immunology 189, 59–66. Finlay, B.B. and Cossart, P. (1997) Exploitation of mammalian host cell functions by bacterial pathogens. Science 276, 718–725. Fraser, C.M., Casjens, S., Huang, W.M., Sutton, G.G., Clayton, R., Lathigra, R., White, O., Ketchum, K.A., Dodson, R., Hickey, E.K., Gwinn, M., Dougherty, B., Tomb, J.F., Fleischmann, R.D., Richardson, D., Peterson, J., Kerlavage, A.R., Quackenbush, J., Salzberg, S., Hanson, M., van Vugt, R., Palmer, N., Adams, M.D., Gocayne, J., Weidman, J., Utterback, T., Watthey, L., McDonald, L., Artiach, P., Bowman, C., Garland, S., Fujii, C., Cotton, M.D., Horst, K., Roberts, K., Hatch, B., Smith, H.O. and Venter, J.C. (1997) Genomic sequence of a Lyme disease spirochaete, Borrelia burgdorferi. Nature 390, 580–586. Fuchs, H., Wallich, R., Simon, M.M. and Kramer, M.D. (1994) The outer surface protein A of the spirochete Borrelia burgdorferi is a plasmin(ogen) receptor. Proceedings of the National Academy of Sciences, USA 91, 12594–12598. Fuchs, H., Simon, M.M., Wallich, R., Bechtel, M. and Kramer, M.D. (1996) Borrelia burgdorferi induces secretion of pro-urokinase-type plasminogen activator by human monocytes. Infection and Immunity 64, 4307–4312. Gern, L., Schaible, U.E. and Simon, M.M. (1993) Mode of inoculation of the Lyme disease agent Borrelia burgdorferi influences infection and immune responses in inbred strains of mice. Journal of Infectious Diseases 167, 971–975. Gern, L., Siegenthaler, M.C., Hu, M., Leuba-Garcia, S., Humair, P.F. and Moret, J. (1994) Borrelia burgdorferi in rodents (Apodemus flavicollis and A. sylvaticus): duration and enhancement of infectivity for Ixodes ricinus ticks. European Journal of Epidemiology 10, 75–80.

142

K. Kurtenbach et al.

Gern, L., Estrada-Pena, A., Frandsen, F., Gray, J.S., Jaenson, T.G.T., Jongejan, F., Kahl, O., Korenberg, E., Mehl, R. and Nuttall, P.A. (1998) European reservoir hosts of Borrelia burgdorferi sensu lato. Zentralblatt für Bakteriologie 287, 196–204. Gilmore, R.D. and Piesman, J. (2000) Inhibition of Borrelia burgdorferi migration from the midgut to the salivary glands following feeding by ticks on OspCimmunized mice. Infection and Immunity 68, 411–414. Golde, W.T., Burkot, T.R., Sviat, S., Keen, M.G., Mayer, L.W., Johnson, B.J.B. and Piesman. J. (1993) The major histocompatibility complex-restricted response of recombinant inbred strains of mice to natural tick transmission of Borrelia burgdorferi. Journal of Experimental Medicine 177, 9–17. Grab, D.J., Givens, C. and Kennedy, R. (1998) Fibronectin-binding activity in Borrelia burgdorferi. Biochimica et Biophysica Acta 1407, 135–145. Grab, D.J., Lanners, H.-N., Martin, L.N., Chesney, J., Cai, C., Adkisson, H.D. and Bucala, R. (1999) Interaction of Borrelia burgdorferi with peripheral blood fibrocytes, antigen-presenting cells with the potential for connective tissue targeting. Molecular Medicine 5, 46–54. Guo, B.P., Norris, S.J., Rosenberg, L.C. and Hook, M. (1995) Adherence of Borrelia burgdorferi to the proteoglycan decorin. Infection and Immunity 63, 3467–3472. Gylfe, Å., Olsen, B., Strasevicius, D., Ras, N.M., Weihe, P., Noppa, L., Ostberg, Y., Baranton, G. and Bergström, S. (1999) Isolation of Lyme disease Borrelia from puffins (Fratercula arctica) and seabird ticks (Ixodes uriae) on the Faroe Islands. Journal of Clinical Microbiology 37, 890–896. Gylfe, Å., Bergström, S., Lunstrom, J. and Olsen, B. (2000) Epidemiology – reactivation of Borrelia infection in birds. Nature 403, 724–725.

Hellwage, J., Meri, T., Heikkilae, T., Alitalo, A., Panelius, J., Lahdenne, P., Seppaelae, I.J.T. and Meri, S. (2001) The complement regulator factor H binds to the surface protein OspE of Borrelia burgdorferi. Journal of Biological Chemistry 276, 8427–8435. Hodzic, E., Febg, S. and Barthold, S.W. (2000) Stability of Borrelia burgdorferi outer surface protein C under immune selection pressure. Journal of Infectious Diseases 181, 750–753. Hofmeister, E.K., Ellis, B.A., Glass, G.E. and Childs, J.E. (1999a) Longitudinal study of infection with Borrelia burgdorferi in a population of Peromyscus leucopus at a Lyme disease-enzootic site in Maryland. American Journal of Tropical Medicine and Hygiene 60, 598–609. Hofmeister, E.K., Glass, G.E., Childs, J.E. and Persing, D.H. (1999b) Population dynamics of a naturally occurring heterogeneous mixture of Borrelia burgdorferi clones. Infection and Immunity 67, 5709–5716. Hu, C.M., Wilske, B., Fingerle, V., Lobet, Y. and Gern, L. (2001) Transmission of Borrelia garinii OspA serotype 4 to BALB/c mice by Ixodes ricinus ticks collected in the field. Journal of Clinical Microbiology 39, 1169–1171. Humair, P.F., Peter, O., Wallich, R. and Gern, L. (1995) Strain variation of Lyme disease spirochetes isolated from Ixodes ricinus ticks and rodents collected in two endemic areas in Switzerland. Journal of Medical Entomology 32, 433–438. Humair, P.F., Postic, D., Wallich, R. and Gern, L. (1998) An avian reservoir (Turdus merula) of the Lyme borreliosis spirochetes. Zentralblatt für Bakteriologie 287, 521–538. Humair, P.F., Rais, O. and Gern, L. (1999) Transmission of Borrelia afzelii from Apodemus mice and Clethrionomys voles to Ixodes ricinus ticks, differential transmission pattern and overwintering maintenance. Parasitology 118, 33–42.

B. burgdorferi in the Vertebrate Host

Jauris-Heipke, S., Liegl, G., Preac-Mursic, V., Roessler, D., Schwab, E., Soutschek, E., Will, G. and Wilske, B. (1995) Molecular analysis of genes encoding outer surface protein C (OspC) of Borrelia burgdorferi sensu lato: relationship to ospA genotype and evidence of lateral gene exchange of ospC. Journal of Clinical Microbiology 33, 1860–1866. Kirstein, F., Rijpkema, S., Molkenboer, M. and Gray, J.S. (1997) Local variations in the distribution of Borrelia burgdorferi sensu lato genomospecies in Ixodes ricinus ticks. Applied and Environmental Microbiology 3, 1102–1106. Klempner, M.S., Noring, R. and Rogers, R.A. (1993) Invasion of human skin fibroblasts by the Lyme disease spirochaete, Borrelia burgdorferi. Journal of Infectious Diseases 167, 1074–1081. Knigge, H., Simon, M.M., Meuer, S.C., Kramer, M.D. and Wallich, R. (1996) The outer surface lipoprotein OspA of Borrelia burgdorferi provides co-stimulatory signals to normal human peripheral CD4+ and CD8+ T lymphocytes. European Journal of Immunology 26, 2299–2303. Kraiczy, P., Skerka, C., Kirschfink, M., Brade, V. and Zipfel, P.F. (2001a) Immune evasion of Borrelia burgdorferi by acquisition of human complement regulators FHL-1/reconectin and factor H. European Journal of Immunology 31, 1674–1684. Kraiczy, P., Skerka, C., Kirschfink, M., Zipfel, P.F. and Brade, V. (2001b) Mechanism of complement resistance of pathogenic Borrelia burgdorferi isolates. International Immunopharmacology 1, 393–401. Kuo, M.M., Lane, R.S. and Giclas, P.C. (2000) A comparative study of mammalian and reptilian alternative pathway of complement-mediated killing of the Lyme disease spirochete (Borrelia burgdorferi). Journal of Parasitology 86, 1223–1228.

143

Kurtenbach, K., Dizij, A., Seitz, H.-M., Margos, G., Moter, S.E., Kramer, M.D., Wallich, R., Schaible, U.E. and Simon, M.M. (1994) Differential immune responses to Borrelia burgdorferi in European wild rodent species influence spirochete transmission to Ixodes ricinus L. (Acari: Ixodidae). Infection and Immunity 62, 5344–5352. Kurtenbach, K., Kampen, H., Dizij, A., Arndt, S., Seitz, H.-M., Schaible, U.E. and Simon, M.M. (1995) Infestation of rodents with larval Ixodes ricinus L. (Acari: Ixodidae) is an important factor in the transmission cycle of Borrelia burgdorferi s.l. in German woodlands. Journal of Medical Entomology 32, 807–817. Kurtenbach, K., Dizij, A., Voet, P., Hauser, P. and Simon, M.M. (1997) Vaccination of natural reservoir hosts with recombinant lipidated OspA induces a transmission-blocking immunity against Lyme disease spirochaetes associated with high levels of LA-2 equivalent antibodies. Vaccine 15, 1670–1674. Kurtenbach, K., Carey, D., Hoodless, A., Nuttall, P.A. and Randolph, S.E. (1998a) Competence of pheasants as reservoirs for Lyme disease spirochetes. Journal of Medical Entomology 35, 77–81. Kurtenbach, K., Peacey, M., Rijpkema, S.G.T., Hoodless, A.N., Nuttall, P.A. and Randolph, S.E. (1998b) Differential transmission of the genospecies of Borrelia burgdorferi sensu lato by game birds and small rodents in England. Applied and Environmental Microbiology 64, 1169–1174. Kurtenbach, K., Sewell, H.-S., Ogden, N.H., Randolph, S.E. and Nuttall, P.A. (1998c) Serum complement sensitivity as a key factor in Lyme disease ecology. Infection and Immunity 66, 1248–1251. Kurtenbach, K., de Michelis, S., Sewell, H.-S., Etti, S., Schäfer, M., Hails, R., Collares-Pereira, M., Haninçová, K.,

144

K. Kurtenbach et al.

Labuda, M., Bormane, A. and Donaghy, M. (2001) Distinct combinations of Borrelia burgdorferi sensu lato genospecies found in individual questing ticks from Europe. Applied and Environmental Microbiology 67, 4926– 4929. Lachmann, P.J. (1986) A common form of killing. Nature 321, 560. Lane, R.S. and Quistad, G.B. (1998) Borreliacidal factor in the blood of the western fence lizard (Sceloporus occidentalis). Journal of Parasitology 84, 29–34. Lebet, N. and Gern, L. (1994) Histological examination of Borrelia burgdorferi infections in unfed Ixodes ricinus nymphs. Experimental and Applied Acarology 18, 177–183. Le Fleche, A., Postic, D., Girardet, K., Peter, O. and Baranton, G. (1997) Characterization of Borrelia lusitaniae sp. nov. by 16S ribosomal DNA sequence analysis. International Journal of Systematic Bacteriology 47, 921–925. Ma, Y. and Weiss, J.J. (1993) Borrelia burgdorferi outer surface lipoproteins OspA and OspB possess B-cell mitogenic and cytokine-stimulatory properties. Infection and Immunity 61, 3843–3853. Marconi, R.T., Hohenberger, S., JaurisHeipke, S., Schulte-Spechtel, U., Lavoie, C.P., Rößler, D. and Wilske, B. (1999) Genetic analysis of Borrelia garinii OspA serotype 4 strains associated with neuroborreliosis: evidence for genetic homogeneity. Journal Clinical Microbiology 37, 3965–3970. Marti-Ras, N., Postic, D., Foretz, M. and Baranton, G. (1997) Borrelia burgdorferi sensu stricto, a bacterial species ‘made in the U.S.A.’? International Journal of Systematic Bacteriology 47, 1112–1117. Masuzawa, T., Iwaki, A., Sato, Y., Miyamoto, K., Korenberg, E.I. and Yanagihara, Y. (1997) Genetic diversity of Borrelia burgdorferi sensu lato isolated in far eastern Russia. Microbiology and Immunology 41, 595–600. Masuzawa, T., Takada, N., Kudeken, M.,

Fukui, T., Yano, Y., Ishiguro, F., Kawamura, Y., Imai, Y. and Ezaki, T. (2001) Borrelia sinica sp. nov., a Lyme disease-related Borrelia species isolated in China. International Journal of Systematic and Evolutionary Microbiology 51, 1817–1824. Mather, T.N., Wilson, M.L., Moore, S.I., Ribeiro, J.M. and Spielman, A. (1989) Comparing the relative potential of rodents as reservoirs of the Lyme disease spirochete (Borrelia burgdorferi). American Journal of Epidemiology 130, 143–150. Matuschka, F.R., Heiler, M., Eiffert, H., Fischer, P., Lotter, H. and Spielman, A. (1993) Diversionary role of hoofed game in the transmission of Lyme disease spirochetes. American Journal of Tropical Medicine and Hygiene 48, 693–699. Maupin, G.O., Gage, K.L., Piesman, J., Montenieri, S.L., Sviat, L., VanderZanden, C.M., Happ, M., Dolan, M. and Johnson, B.J.B. (1994) Discovery of an enzootic cycle of Borrelia burgdorferi in Neotoma mexicana and Ixodes spinipalpis from northern Colorado, an area where Lyme disease is nonendemic. Journal of Infectious Diseases 170, 636–643. de Michelis, S., Sewell, H.-S., CollaresPereira, M., Santos-Reis, M., Schouls, L.M., Benes, V., Holmes, E.C. and Kurtenbach, K. (2000) Genetic diversity of Borrelia burgdorferi sensu lato in ticks from mainland Portugal. Journal of Clinical Microbiology 38, 2128–2133. Modolell, M., Schaible, U.E., Rittig, M. and Simon, M.M. (1994) Killing of Borrelia burgdorferi by macrophages is dependent on oxygen radicals and nitric oxide and can be enhanced by antibodies to outer surface proteins of the spirochete. Immunology Letters 40, 139–146. Montgomery, R.R. and Malavista, S.E. (1996) Entry of Borrelia burgdorferi into macrophages is end-on and leads to degradation in lysosomes. Infection and Immunity 64, 2867–2972. Museteanu, C., Schaible, U.E., Stehle, T., Kramer, M.D. and Simon, M.M. (1991) Myositis in mice inoculated with Borrelia

B. burgdorferi in the Vertebrate Host

burgdorferi. American Journal of Pathology 139, 1267–1271. Nakao, M., Miyamoto, K. and Fukunaga, M. (1994a) Lyme disease spirochetes in Japan: enzootic transmission cycles in birds, rodents, and Ixodes persulcatus ticks. Journal of Infectious Diseases 170, 878–882. Nakao, M., Miyamoto, K. and Fukunaga, M. (1994b) Borrelia japonica in nature: genotypic identification of spirochetes isolated from Japanese small mammals. Microbiology and Immunology 38, 805–808. Nelson, D.R., Rooney, S., Miller, N.J. and Mather, T.N. (2000) Complementmediated killing of Borrelia burgdorferi by non-immune sera from sika deer. Journal of Parasitology 86, 1232–1238. Nordstrand, A., Barbour, A.G. and Bergström, S. (2000) Borrelia pathogenesis research in the post-genomic and post-vaccine era. Current Opinion in Microbiology 3, 86–92. Ohnishi, J., Piesman, J. and de Silva, A. (2001) Antigenic and genetic heterogeneity of Borrelia burgdorferi populations transmitted by ticks. Proceedings of the National Academy of Sciences, USA 98, 670–675. Oliver, J.H. (1996) Lyme borreliosis in the southern United States: a review. Journal of Parasitology 82, 926–935. Olsen, B., Jaenson, T.G.T., Noppa, L., Bunikis, J. and Bergström, S.A. (1993) A Lyme borreliosis cycle in seabirds and Ixodes uriae ticks. Nature 362, 340–342. Olsen, B., Duffy, D.C., Jaenson, T.G.T., Gylfe, A., Bonnedahl, J. and Bergström, S.A. (1995a) Transhemispheric exchange of Lyme disease spirochaetes by seabirds. Journal of Clinical Microbiology 33, 3270–3274. Olsen, B., Jaenson, T.G.T. and Bergström, S.A. (1995b) Prevalence of Borrelia burgdorferi sensu lato-infected ticks on migrating birds. Applied and Environmental Microbiology 61, 3082–3087. Olsen, B., Gylfe, A. and Bergström, S.A. (1996) Canary finches (Serinus canaria) as an avian infection model for Lyme

145

borreliosis. Microbial Pathogenesis 20, 319–324. Oteo, J.A., Backenson, P.B., Vitucia, M.M., Garcia Monco, J.C., Rodriguez, L., Escudero, R. and Anda, P. (1998) Use of C3H/He Lyme disease mouse model for the recovery of a Spanish isolate of Borrelia garinii from erythema migrans lesions. Research in Microbiology 149, 39–46. Pal, U., de Silva, A.M., Montgomery, R.R., Fish, D., Anguita, J., Anderson, J.F., Lobet, Y. and Fikrig, E. (2000) Attachment of Borrelia burgdorferi within Ixodes scapularis mediated by outer surface protein A. Journal of Clinical Investigation 106, 561–569. Papatheodorou, V. and Brossard, M. (1987) C-3 levels in the sera of rabbits infested and reinfested with Ixodes ricinus L. and in midguts of fed ticks. Experimental and Applied Acarology 3, 53–59. Parveen, N., Robbins, D. and Leong, J.M. (1999) Strain variation in glycosaminoglycan recognition influences cell-type specific binding by Lyme disease spirochaetes. Infection and Immunity 67, 1743–1749. Pennington, P.M., Allred, C.D., West, C.S., Alvarez, R. and Barbour, A.G. (1997) Arthritis severity and spirochete burden are determined by serotype in the Borrelia turicatae–mouse model of Lyme disease. Infection and Immunity 65, 285–292. Philipp, M.T., Aydintug, M.K., Bohm, R.P. Jr, Cogswell, F.B., Dennis, V.A., Lanners, H.N., Lowrie, R.C. Jr, Roberts, E.D., Conway, M.D., Karacorlu, M. et al. (1993) Early and early-disseminated phases of Lyme disease in the rhesus monkey: a model for infections in humans. Infection and Immunity 61, 3047–3059. Piesman, J. (1988) Intensity and duration of Borrelia burgdorferi and Babesia microti infectivity in rodent hosts. International Journal of Parasitology 18, 687–689.

146

K. Kurtenbach et al.

Piesman, J. (1991) Experimental acquisition of the Lyme disease spirochete, Borrelia burgdorferi, by larval Ixodes dammini (Acari: Ixodidae) during partial blood meals. Journal of Medical Entomology 28, 259–262. Piesman, J. (1993a) Dynamics of Borrelia burgdorferi transmission by nymphal Ixodes dammini ticks. Journal of Infectious Diseases 167, 1082–1085. Piesman, J. (1993b) Standard system for infecting ticks (Acari: Ixodidae) with the Lyme disease spirochete, Borrelia burgdorferi. Journal of Medical Entomology 30, 199–203. Piesman, J., Mather, T.N., Sinsky, R.J. and Spielman, A. (1987) Duration of tick attachment and Borrelia burgdorferi transmission. Journal of Clinical Microbiology 25, 557–558. Piesman, J., Oliver, J.O. and Sinsky, R.J. (1990) Growth kinetics of the Lyme disease spirochete (Borrelia burgdorferi) in vector ticks (Ixodes dammini). American Journal of Tropical Medicine and Hygiene 42, 352–357. Piesman, J., Maupin, G.O., Campos, E.G. and Happ, C.M. (1991) Duration of adult female Ixodes dammini attachment and transmission of Borrelia burgdorferi, with description of a needle aspiration isolation method. Journal of Infectious Diseases 163, 895–897. Piesman, J., Dolan, M.C., Schriefer, M.E. and Burkot, T.R. (1996) Ability of experimentally infected chickens to infect ticks with the Lyme disease spirochete, Borrelia burgdorferi. American Journal of Tropical Medicine and Hygiene 54, 294–298. Postic, D., Assous, M., Grimont, P.A.D. and Baranton, G. (1994) Diversity of Borrelia burgdorferi sensu lato evidenced by restriction fragment length polymorphism of rrf (5S)–rrl (23S) intergenic spacer amplicons. International Journal of Systematic Bacteriology 44, 743–752. Postic, D., Marti-Ras, N., Lane, R.S., Hendson, M. and Baranton, G. (1998)

Expanded diversity among Californian Borrelia isolates and description of Borrelia bissettii sp. nov. (formerly Borrelia group DN127). Journal of Clinical Microbiology 36, 3497–3504. Preac-Mursic, V., Patsouris, E., Wilske, B., Reinhardt, S., Gos, B. and Mehraein, P. (1990) Persistence of Borrelia burgdorferi and histopathological alterations in experimentally infected animals: comparison with histopathological findings in human Lyme disease. Infection 18, 332–341. Qui, W.-G., Bosler, E., Campbell, J.R., Ugine, G.D., Wang, L.-N., Luft, B.J. and Dykhuizen, D.E. (1997) A population genetic study of Borrelia burgdorferi sensu stricto from eastern Long Island, New York, suggested frequencydependent selection, gene flow and host adaptation. Hereditas 127, 203–216. Radolf, J.D., Arndt, L.L., Akins, D.R., Curetty, L.L., Levi, M.E., Davis, L.S. and Norgard, M.V. (1995) Treponema pallidum and Borrelia burgdorferi lipoproteins and synthetic lipoproteins activate monocytes/macrophages. Journal of Immunology 154, 2866–2877. Rand, P.W., Lacombe, E.H., Smith, R.P. and Ficker, J. (1998) Participation of birds (Aves) in the emergence of Lyme disease in Southern Maine. Journal of Medical Entomology 35, 270–276. Randolph, S.E. (1998) Ticks are not insects: consequences of contrasting vector biology for transmission potential. Parasitology Today 14, 186–192. Randolph, S.E. and Craine, N.G.A. (1995) A general framework for comparative quantitative studies on the transmission of tick-borne diseases using Lyme borreliosis in Europe as an example. Journal of Medical Entomology 32, 765–777. Richter, D., Spielman, A., Komar, N. and Matuschka, F.R. (2000) Competence of American robins as reservoir hosts for Lyme disease spirochetes. Emerging Infectious Diseases 6, 133–138.

B. burgdorferi in the Vertebrate Host

Rittig, M.G., Jagoda, J.C., Wilske, B., Murgia, R., Cinco, M., Repp, R., Burmester, G.R. and Krause, R. (1998) Coiling phagocytosis discriminates between different spirochetes and is enhanced by phorbol myristate acetate and granulocyte–macrophage colonystimulating factor. Infection and Immunity 66, 627–635. Roehrig, J.T., Piesman, J., Hunt, A.R., Keen, M.G., Happ, C.M. and Johnson, B.J.B. (1992) The hamster immune response to tick-transmitted Borrelia burgdorferi differs from the response to needle-inoculated, cultured organisms. Journal of Immunology 149, 3648–3653. Ryan, J.R., Levine, J.F., Apperson, C.S., Lubke, L., Wirtz, R.A., Spears, P.A. and Orndorff, P.E. (1998) An experimental chain of infection reveals that distinct Borrelia burgdorferi populations are selected in arthropod and mammalian hosts. Molecular Microbiology 30, 365–379. Schaible, U.E., Kramer, M.D., Museteanu, C., Zimmer, G., Mossmann, H. and Simon, M.M. (1989) The severe combined immunodeficiency (scid) mouse. A laboratory model for the analysis of Lyme arthritis and carditis. Journal of Experimental Medicine 170, 1427–1432. Schaible, U.E., Gern, L., Wallich, R., Kramer, M.D., Prester, M. and Simon, M.M. (1993) Distinct patterns of protective antibodies are generated against Borrelia burgdorferi in mice experimentally inoculated with high and low doses of antigen. Immunology Letters 36, 219–226. Schaible, U.E., Vestweber, D., Butcher, E.G., Stehle, T. and Simon, M.M. (1994) Expression of endothelial cell adhesion molecules in joints and heart during Borrelia burgdorferi infection of mice. Cell Adhesion Communication 2, 465–479. Schmitz, J.L., Hejke, A., England, D.M., Callister, S.M. and Schell, R.F. (1989) Induction of Lyme arthritis in LSH hamsters. American Journal of Pathology 134, 1113–1123.

147

Schneider, B.S., Zeidner, N.S., Burkot, T.R., Maupin, G.O. and Piesman, J. (2000) Borrelia isolates in Northern Colorado identified as Borrelia bissettii. Journal of Clinical Microbiology 38, 3103–3105. Schwan, T.G., Burgdorfer, W. and Garon, C.F. (1988) Changes in infectivity and plasmid profile of the Lyme disease spirochete, Borrelia burgdorferi, as a result of in vitro cultivation. Infection and Immunity 56, 1831–1836. Schwan, T.G., Piesman, J., Golde, W.T., Dolan, M.C. and Rosa, P.A. (1995) Induction of an outer surface protein on Borrelia burgdorferi during tick feeding. Proceedings of the National Academy of Sciences, USA 92, 2909–2913. Seinost, G., Dykhuizen, D.E., Datwyler, R.J., Golde, W.T., Dunn, J.J., Wang, L.-N., Wormser, G.P., Schriefer, M.E. and Luft, B.J. (1999) Four clones of Borrelia burgdorferi sensu stricto cause invasive infections in humans. Infection and Immunity 67, 3518–3524. Sellati, T.J., Abrescia, L.D., Radolf, J.D. and Furie, M.B. (1996) Outer surface lipoproteins of Borrelia burgdorferi activate vascular endothelium in vitro. Infection and Immunity 64, 3170–3180. Shih, C.M., Pollack, R.J., Telford, S.R. III and Spielman, A. (1992) Delayed dissemination of Lyme disease spirochetes from the site of deposition in the skin of mice. Journal of Infectious Diseases 166, 827–831. Siebers, A., Zhong, W., Wallich, R. and Simon, M.M. (1999) Loss of pathogenic potential after cloning of the low-passage Borrelia burgdorferi ZS7 tick isolate, a cautionary note. Medical Microbiology and Immunology 188, 125–130. de Silva, A.M. and Fikrig, E. (1997) Borrelia burgdorferi genes selectively expressed in ticks and mammals. Parasitology Today 13, 267–270. de Silva, A.M., Zeidner, N.S., Zhang, Y., Dolan, M.C., Piesman, J. and Fikrig, E. (1999) Influence of outer surface protein

148

K. Kurtenbach et al.

A antibody on Borrelia burgdorferi within feeding ticks. Infection and Immunity 67, 30–35. Sim, R.B. and Dodds, A.W. (1997) The complement system: an introduction. In: Dodds, A.W. and Sim, R.B. (eds) Complement: a Practical Approach. IRL Press, Oxford, pp. 1–18. Simon, M.M., Kramer, M.D., Wallich, R. and Schaible, U.E. (1993) Lyme arthritis: pathogenic principles emerging from studies in man and mouse. In: Panaysi, G.S. (ed.) Immunology of Connective Tissue Diseases. Kluwer Academic Publishers, Dordrecht, The Netherlands, pp. 205–229. Simon, M.M., Nerz, G., Kramer, M.D., Hurtenbach, U., Schaible, U.E. and Wallich, R. (1995) The outer surface lipoprotein A of Borrelia burgdorferi provides direct and indirect augmenting/ co-stimulatory signals for the activation of CD4+ and CD8+ T cells. Immunology Letters 45, 137–142. Sonnensyn, S.W., Manivel, J.C., Johnson, R.C. and Goodman, J.L. (1993) A guinea pig model for Lyme disease. Infection and Immunity 61, 4777–4784. Sprenger, H., Krause, A., Kaufmann, A., Priem, S., Fabian, D., Burmester, G.R., Gemsa, D. and Rittig, M.G. (1997) Borrelia burgdorferi induces chemokines in human monocytes. Infection and Immunity 65, 4384–4388. Stevenson, B., Casjens, S. and Rosa, P. (1998) Evidence of past recombination events among genes encoding the Erp antigens of Borrelia burgdorferi. Microbiology 144, 1869–1879. Sung, S.Y., Lavoie, C.P., Carlyon, J.A. and Marconi, R.T. (1998) Genetic divergence and evolutionary instability of ospE-related members of the upstream homology box gene family in Borrelia burgdorferi sensu lato complex isolates. Infection and Immunity 66, 4656–4668. Telford, S.R. III, Mather, T.N., Moore, S.I., Wilson, M.L. and Spielman, A. (1988) Incompetence of deer as reser-

voirs of the Lyme disease spirochete. American Journal Tropical Medicine and Hygiene 39, 105–109. Thomas, V., Anguita, J., Samanta, S., Rosa, P.A., Stewart, P., Barthold, S.W. and Fikrig, E. (2001) Dissociation of infectivity and pathogenicity in Borrelia burgdorferi. Infection and Immunity 69, 3507–3509. Valenzuela, J.G., Charlab, R., Mather, T.N. and Ribeiro, J.M.C. (2000) Purification, cloning, and expression of a novel salivary anticomplement protein from the tick, Ixodes scapularis. Journal of Biological Chemistry 23, 18717–18723. Wallich, R., Brenner, C., Kramer, M.D. and Simon, M.M. (1995) Molecular cloning and immunological characterization of a novel linear-plasmid-encoded gene, pG, of Borrelia burgdorferi expressed only in vivo. Infection and Immunity 63, 3327–3335. Wang, G., van Dam, A.P. and Dankert, J. (1999) Evidence for frequent OspC gene transfer between Borrelia valaisiana sp. nov. and other Lyme disease spirochetes. FEMS Microbiology Letters 177, 289–296. Zhang, J.R., Hardham, J.M., Barbour, A.G. and Norris, S.J. (1997) Antigenic variation in Lyme disease borreliae by promiscuous recombination of VMPlike sequence cassettes. Cell 89, 275–285. Zhong, W., Stehle, T., Museteanu, C., Siebers, A., Gern, L., Kramer, M., Wallich, R. and Simon, M.M. (1997) Therapeutic passive vaccination against chronic Lyme disease in mice. Proceedings of the National Academy of Sciences, USA 94, 12533–12538. Zhong, W., Gern, L., Stehle, T., Museteanu, C., Kramer, M., Wallich, R. and Simon, M.M. (1999) Resolution of experimental and tick-borne Borrelia burgdorferi infection in mice by passive, but not active immunization using recombinant OspC. European Journal of Immunology 29, 946–957.

Ecology of Borrelia burgdorferi sensu lato in Europe

6

Lise Gern and Pierre-François Humair Institute of Zoology, University of Neuchâtel, Emile Argand 11, 2007 Neuchâtel, Switzerland

Introduction Borrelia burgdorferi, the aetiologic agent of the tick-borne zoonosis Lyme borreliosis (LB), circulates between ticks and vertebrate hosts. However, as in the case of most other vector-borne zoonoses, the presence of the pathogen in nature was revealed through clinical manifestations in humans. It took almost 100 years after the first description, in Europe, of clinical manifestations of the disease now known as LB for biologists to discover borreliae in Ixodes scapularis ticks from northeastern USA (Burgdorfer et al., 1982). Thus, the pathogen and the vector of this disease were discovered simultaneously. Similar spirochaetes were observed in Europe shortly thereafter in a closely related tick species, Ixodes ricinus (Burgdorfer et al., 1983). Finally, the spirochaetes were recognized as a new species, B. burgdorferi (Johnson et al., 1984), and as the causative agent of the new clinical entity, Lyme disease or LB. It is now known that LB is the most prevalent tick-borne disease in the northern hemisphere. As in other vector-borne diseases, understanding the ecological relationships that exist between pathogens, vectors and wildlife hosts in LB is essential to the understanding of the epidemiology of the disease in human populations. In addition to the diversity of vector and host species in the large geographical area occupied by B. burgdorferi worldwide, a great diversity of Borrelia species within the complex B. burgdorferi sensu lato (s.l.) has been discovered recently. The transmission and maintenance cycles of B. burgdorferi s.l. in enzootic areas show a complicated picture with regional and local pecularities, and our present view of the ecology of these spirochaetes in Europe is like a collection of temporal and local snapshots, the whole picture being still fragmentary. This chapter is a review of the current knowledge of B. burgdorferi s.l. ecology in Europe, but © 2002 CAB International. Lyme Borreliosis: Biology, Epidemiology and Control (eds J. Gray, O. Kahl, R.S. Lane and G. Stanek)

149

150

L. Gern and P.-F. Humair

because of the relatively few publications from eastern Europe it focuses particularly on western and central regions. The maintenance of B. burgdorferi s.l. requires the interaction of three groups of organisms: the pathogens, vectors and hosts. These three essential components of LB will be described first, before attention is focused on their interrelationships.

The Spirochaetes With the increasing number of isolates of B. burgdorferi obtained from various geographical and biological sources, it became obvious first in Europe and Asia and then also in North America that the spirochaete exhibited phenotypic and genotypic heterogeneity. This resulted in the description of 11 genomic groups or genospecies gathered under the name Borrelia burgdorferi s.l.: B. burgdorferi sensu stricto (s.s.) (Johnson et al., 1984), Borrelia garinii (Baranton et al., 1992), Borrelia afzelii (Canica et al., 1993), Borrelia japonica (Kawabata et al., 1993), Borrelia andersonii (Marconi et al., 1995), Borrelia tanukii and Borrelia turdi (Fukunaga et al., 1996), Borrelia valaisiana (Wang et al., 1997), Borrelia lusitaniae (Le Fleche et al., 1997), Borrelia bissettii (Postic et al., 1998) and Borrelia sinica (Masuzawa et al., 2001). In Europe, five species (B. burgdorferi s.s., B. garinii, B. afzelii, B. valaisiana and B. lusitaniae) have been isolated from ticks and B. burgdorferi s.l. has been reported from 26 countries from Italy to Iceland, and from Portugal to Russia (Hubálek and Halouzka, 1997). Two additional genospecies have been reported on the European continent from patient tissues: B. bissettii, a species present in North America, has been obtained from patients in Slovenia (Picken et al., 1996; Strle et al., 1997), and a novel B. burgdorferi s.l. genospecies has been cultured from an erythema migrans biopsy of a patient who contracted the disease in The Netherlands (Wang et al., 1999). The European vector ticks and natural hosts of these two genospecies have not been determined as yet. In addition to the classification of B. burgdorferi into 11 different species, an intraspecific heterogeneity has been documented with respect to outer surface protein (Osp) A, which is expressed by most B. burgdorferi s.l. spirochaetes (Will et al., 1995). Wilske et al. (1993) developed a comprehensive serotyping system of B. burgdorferi s.l. by using immunoblotting and a variety of monoclonal antibodies against OspA. A strong correlation exists between OspA phenotypes and genospecific grouping of Lyme disease borreliae. This OspA serotyping was confirmed by sequence analysis of the osp genes (Jauris-Heipke et al., 1995; Will et al., 1995). These studies revealed that B. garinii was phenotypically the most heterogeneous genospecies with five different serotypes (serotypes 3–7).

Vectors and Carriers In Europe, B. burgdorferi s.l. is mainly transmitted by I. ricinus. Two additional tick species, Ixodes hexagonus and Ixodes uriae, are also vectors and maintain

Ecology of B. burgdorferi in Europe

151

B. burgdorferi s.l. in secondary transmission cycles without necessarily involving I. ricinus. The sheep tick or castor bean tick, I. ricinus, the principal vector of B. burgdorferi s.l., is the most common tick species in Europe. I. ricinus occurs from the Atlantic coast to as far east as 50–60° longitude in central Asia and from Iceland to as far south as the Atlas Mountains of North Africa (Hillyard, 1996). The eastern distribution of I. ricinus overlaps the western distribution of the related Ixodes persulcatus (Fig. 6.1). The habitat of I. ricinus is situated in the leaf litter and the low strata vegetation of temperate deciduous woodlands and mixed forests. In areas with high rainfall, I. ricinus also occurs in high densities in coniferous forests and in open areas such as grassland (Gray et al., 1998). I. ricinus has particular requirements regarding the humidity of its habitat and a relative humidity (RH) value of > 80% is important for the activity and survival of free-living I. ricinus. A low

ICELAND

I. persulcatus

Norwegian Sea

SWEDEN

I. ricinus

FINLAND Gulf of Bothnia

NORWAY

North Atlantic Ocean

RUSSIA ESTONIA

North Sea LATVIA

DENMARK

Baltic Sea

IRELAND

LITHUANIA

U. K.

BYELARUS

NETH.

GERMANY

BELGIUM

English Channel

POLAND

UKRAINE LUX. CZECH SLOVAKIA

MOLDOVA

FRANCE

AUSTRIA HUNGARY

SWITZERLAND

Bay of Biscay

ROMANIA

SLOVENIA CROATIA

Black Sea

BOSNIA SERBIA

BULGARIA

ANDORRA

ITALY

PORTUGAL

MONTENEGRO

Adriatic

SPAIN Tyrrhenian Sea

MACEDONIA

ALBANIA

GREECE

Mediterranean Sea

TURKEY Aegean

Ionian Sea CYPRUS

Fig. 6.1. Distribution of Ixodes ricinus and Ixodes persulcatus, the major vectors of Borrelia burgdorferi s.l. in Europe. Note that, in southern areas and in Iceland and northern Scandinavia I. ricinus populations are sparse and scattered.

152

L. Gern and P.-F. Humair

saturation deficit, which is calculated by using values of RH and temperature, can also be closely correlated with optimal conditions for I. ricinus (Randolph and Storey, 1999; Perret et al., 2000). Although studied for more than 50 years, knowledge of the biology of I. ricinus remains incomplete, partly because of the geographical and genetic heterogeneity of this tick species and also because of the diverse habitats it occupies. I. ricinus is an exophilic (or non-nidicolous) tick and the questing tick waits passively on the vegetation for passing hosts. A wide range of vertebrates such as reptiles, birds, small-, medium- and large-sized mammals serve as hosts for I. ricinus (Aeschlimann, 1972). Small mammals, ground-foraging birds and reptiles are common hosts for I. ricinus immatures but not for adults. Medium- to large-sized mammals are parasitized by I. ricinus adults (females) and immatures. Males occasionally take a blood meal but do not engorge. Immature stages and adult females can attach to humans. The life cycle of I. ricinus lasts 2–6 years, typically 3 years, since each stage takes about a year to develop to the next instar (Gray, 1991). The duration of the life cycle can vary from one habitat to another and also regionally, and can be affected by microclimatic factors and host density (Gray, 1991). I. ricinus ticks are active from spring to autumn, generally from February to November and seasonal activity of the three active stages usually shows a bimodal pattern with an activity peak in spring and another one in autumn. Tick activity in spring is usually greater than in autumn, except for larvae, which show the oppposite situation in some areas. Activity peaks can be indistinct and a pseudo-bimodal pattern may occur, as observed in Switzerland (Gigon, 1985; Perret et al., 2000). Besides the classical vector, I. ricinus, two other tick species, I. hexagonus and I. uriae, contribute to the circulation of B. burgdorferi s.l. in Europe. The hedgehog tick, I. hexagonus, is widespread in Europe and is also present in North Africa (Arthur, 1968). The presence of B. burgdorferi s.l. in I. hexagonus was first discovered in tick samples collected from hedgehogs (Liebisch et al., 1989). The vector competence of I. hexagonus for B. burgdorferi s.l. has been demonstrated under both laboratory conditions (Gern et al., 1991; Toutoungi and Gern, 1993) and field conditions (Gern et al., 1997). In contrast to I. ricinus, I. hexagonus is an endophilic (or nidicolous) tick species and its habitat is the nest, burrow or cave of its vertebrate host. The host range of I. hexagonus is therefore more restricted than that of I. ricinus and it feeds primarily on carnivores such as foxes and mustelids, and on hedgehogs, but also, less frequently, on rodents, hares and rabbits (Arthur, 1953; Toutoungi et al., 1991). I. hexagonus has occasionally been collected from birds (Pica pica, Falco tinnunculus) and deer (Capreolus capreolus) (Toutoungi et al., 1991; Hubbard et al., 1998). Domestic animals such as cats, dogs, horses, goats and cows have also been found to be infested (Arthur, 1968; Toutoungi et al., 1991; Bernasconi et al., 1997). I. hexagonus also bites humans but apparently less frequently than I. ricinus (Arthur, 1953; Liebisch and Liebisch, 1996; Liebisch et al., 1998a). The occurrence of I. hexagonus in the urban environment is due to the presence of suitable hosts such as hedgehogs, cats and dogs in gardens and public parks (Gern et al., 1991, 1997).

Ecology of B. burgdorferi in Europe

153

The seabird tick, I. uriae, has been found to be infected by B. garinii (Olsen et al., 1993; Nuttall et al., 1994; Hubbard et al., 1998) and is the third ixodid species implicated in the circulation of B. burgdorferi s.l. in Europe. I. uriae infests more than 50 seabird species living in high latitude areas in both hemispheres (Olsen, 1995). In Europe, I. uriae is present on the coasts of France, Great Britain, Ireland, Denmark, Norway, Sweden and Iceland (Olsen et al., 1995a; Hillyard, 1996). Apparently, I. uriae can infest other avian hosts such as passerines resting in marine habitats (Arthur, 1968). Mammals such as seals, river otters and humans can also occasionally be parasitized by I. uriae (Arthur, 1968; Eley, 1977; Mehl and Traavik, 1983; Hubbard et al., 1998). People engaged in ornithological activity, such as bird-ringing, or in exploitation of colonial seabirds, as in the Faeroe Islands, are likely to be exposed to I. uriae bites (Olsen et al., 1995a; Hubbard et al., 1998; Gylfe et al., 1999). The vector competence of I. uriae for B. burgdorferi s.l. has never been demonstrated under laboratory conditions, but its involvement as a vector of borrelial spirochaetes in transmission cycles in seabird colonies has been shown (Olsen et al., 1993, 1995a). The presence of B. burgdorferi s.l. in vector arthropods results from blood meals taken from infectious vertebrate hosts. In competent vectors, borrelial infection persists in the tick throughout the moulting process to the next instar. This phenomenon, called transstadial transmission, or more appropriately transstadial maintenance, is fundamental to the long-term survival of B. burgdorferi s.l. in nature and its efficiency is one of the factors determining vector competence. Transstadial maintenance has been demonstrated in two European tick vectors, I. ricinus and I. hexagonus (Monin et al., 1989; Gern et al., 1991; Toutoungi and Gern, 1993; Bellet-Edimo, 1997), and apparently occurs also in I. uriae, since unfed I. uriae have been found to be infected (Hubbard et al., 1998) and since a cycle has been described between seabirds and I. uriae (Olsen et al., 1993). B. burgdorferi s.l. is ingested during the tick blood meal, colonizes the tick midgut and usually remains there after the moult. In some cases, spirochaetes may migrate to other tick organs, causing generalized or systemic infections in unfed I. ricinus ticks (Lebet and Gern, 1994; Leuba-Garcia et al., 1994). In female ticks, B. burgdorferi s.l. may be present in ovaries and in the developing embryos (eggs). This mode of infection of the larval progeny through the ovarian tissue of the mother tick is called transovarial or vertical transmission. Experimental transovarial transmission of B. burgdorferi s.s. (isolate Sta14, now identified as NE14) in I. ricinus was successfully demonstrated (Monin et al., 1989). Studies on the efficiency of transovarial transmission showed that this phenomenon rarely occurs in I. ricinus, as only about 1% of field-collected females transmitted B. burgdorferi s.l. to I. ricinus larvae (Bellet-Edimo, 1997). However, when transovarial transmission does occur, the prevalence of infection in I. ricinus eggs or subsequent larvae is high (43–100%) (Burgdorfer et al., 1983; Bellet-Edimo, 1997). Transovarial transmission of B. burgdorferi s.s. (isolate B31) also occurs in I. hexagonus and has been shown experimentally (Gern et al., 1991). In contrast to I. ricinus, many more I. hexagonus females can transmit B. burgdorferi s.l. to their progeny, but infection rates within batches of infected progenies are lower than

154

L. Gern and P.-F. Humair

in I. ricinus. This discrepancy might be related to tick biological and behavioural features: in the case of the endophilic I. hexagonus, a high prevalence of infection in larvae is less important than in the case of the exophilic I. ricinus, because of closer contacts in a confined environment between endophilic ticks and hosts. However, the prevalence of inherited infection in field-derived I. hexagonus larvae remains unknown. At present, no data are available concerning the existence of transovarial transmission of B. burgdorferi s.l. in I. uriae. B. burgdorferi s.l. is reportedly present in all I. ricinus populations examined so far in Europe (Hubálek and Halouzka, 1998). Consequently, the infection risk for humans seems to be present in any I. ricinus habitat. The infection prevalence of questing I. ricinus ticks in Europe averages 1.9% for larvae (range 0–11%), 10.8% for nymphs (range 2–43%) and 17.4% for adults (range 3–75%) (Hubálek and Halouzka, 1998; de Michelis et al., 2000) and varies geographically and also according to the methods used for the detection of B. burgdorferi s.l. in unfed I. ricinus ticks. Cultivation in BSK medium seems to be the least sensitive method compared with microscopy (dark-field, phase-contrast or immunofluorescence) or PCR (Hubálek and Halouzka, 1998; Gern et al., 1999). The prevalence of B. burgdorferi s.l. in unfed I. hexagonus is unknown. Concerning I. uriae, the prevalence of B. garinii was reported by Olsen (1995), who detected borreliae in 1/100 unengorged ticks, and by Hubbard et al. (1998), who detected infection in 13/13 field-collected engorged or semi-engorged museum specimens. Given the biology of I. hexagonus and I. uriae, inspection of host nests is needed for the collection of additional data for unfed free-living ticks of these two species. As already mentioned, five genospecies of B. burgdorferi s.l. (B. burgdorferi s.s., B. garinii, B. afzelii, B. valaisiana and B. lusitaniae) are present in European ixodid vectors. The three pathogenic species (B. burgdorferi s.s., B. garinii and B. afzelii) have been frequently isolated from or detected in Ixodes spp. ticks, whereas B. valaisiana and B. lusitaniae have rarely been reported. Since the discovery of B. burgdorferi in I. ricinus ticks, valuable data concerning the occurrence of B. burgdorferi s.l. genospecies have been collected throughout Europe. However, it is difficult to draw a precise map of the distribution and frequency of the different B. burgdorferi s.l. genospecies on the European continent until more data have been collected. Current knowledge shows that B. garinii is the most frequent genospecies found in ticks, followed by B. afzelii and B. burgdorferi s.s. (Hubálek and Halouzka, 1997; Saint Girons et al., 1998). B. garinii and B. afzelii occur throughout the European continent from the Atlantic coasts to the Ural mountains. In contrast, the frequency of B. burgdorferi s.s. seems to decrease from west to east (Saint Girons et al., 1998) and appears to be absent in I. persulcatus. B. valaisiana and B. lusitaniae appear to be far less common and have been reported from certain European countries only. B. valaisiana occurs in Switzerland (Péter and Bretz, 1992; Péter et al., 1995; Humair et al., 1998), The Netherlands (Rijpkema et al., 1995), Germany (Liebisch et al., 1998a), Croatia (Rijpkema et al., 1996), Great Britain (Cutler et al., 1989; Kurtenbach et al., 1998b), Ireland (Kirstein et al., 1997), Italy (Cinco et al., 1998) and Slovakia (Gern et al., 1999).

Ecology of B. burgdorferi in Europe

155

B. lusitaniae has been reported in Portugal (Nuncio et al., 1993; de Michelis et al., 2000), Spain (Escudero et al., 2000), Slovakia (Gern et al., 1999), Belorussia, the Czech Republic (Le Fleche et al., 1997), Moldavia, Ukraine (Postic et al., 1997) and interestingly in North Africa (Zhioua et al., 1999; Younsi et al., 2001). It is true that B. valaisiana and B. lusitaniae have been described only recently but this cannot explain their lower frequency compared with pathogenic Borrelia spp. Very few isolates of B. valaisiana and B. lusitaniae have been obtained to date, and cultivation in the usual BSK medium seems to be problematic for the isolation of these two genomic groups. However, PCR-based studies have demonstrated a high prevalence of B. valaisiana in unfed I. ricinus ticks in Ireland (Kirstein et al., 1997) and of B. lusitaniae in ticks in Portugal (de Michelis et al., 2000). The question remains whether the DNA detected reflects the presence of viable borreliae. Further studies are needed to determine if B. valaisiana and B. lusitaniae are indeed patchily distributed in Europe or if they have been relatively undetected previously. Infections by multiple B. burgdorferi s.l. genospecies have been observed in ticks in many parts of Europe, including The Netherlands (Rijpkema et al., 1995), Croatia (Rijpkema et al., 1996), Switzerland (Leuba-Garcia et al., 1994), France (Pichon et al., 1995), Belgium (Misonne et al., 1998), Estonia, Kirghizia, Moldavia, Russia and Ukraine (Postic et al., 1997), Ireland (Kirstein et al., 1997), Italy (Cinco et al., 1998) and Germany (Liebisch et al., 1998a; Hu et al., 2001). Different combinations of mixed infection with two or three genospecies have been detected. B. burgdorferi s.s. and B. lusitaniae are the two Borrelia spp. least involved in mixed infections. In most cases, mixed infections in ticks have been detected using PCR methods. Detection of mixed infections using cultivation might be difficult because one genospecies may overgrow another. Certain tick or insect species found to be infected, but without evidence of vector competence, are considered to be carriers (Table 6.1). Transstadial transmission appears to occur in some tick carrier species since unfed ticks have been found infected by borreliae. It should be noted that spirochaetes in these insect and tick carriers have not always been clearly characterized as B. burgdorferi s.l.

Vertebrate Hosts B. burgdorferi s.l. is maintained in natural cycles involving not only vectors but also wild vertebrates. The vertebrate hosts that are required to maintain tick populations may also act as reservoir hosts for tick-transmitted pathogens. Reservoir hosts as defined by Kahl et al. (Chapter 2) have mainly been identified in Europe by xenodiagnosis and by comparison of infection rates in questing ticks with those in ticks removed from hosts. These reservoir hosts are among the large spectrum of different host species of the three competent tick vectors, I. ricinus, I. hexagonus and I. uriae. I. ricinus feeds on the largest variety of hosts on the European continent, infesting more than 300 different species including mammals, birds and reptiles (Anderson,

156

L. Gern and P.-F. Humair

Table 6.1. Arthropod carrier species of Borrelia burgdorferi sensu lato in Europe. Species

Countries

ACARI Ixodidae Ixodes trianguliceps

France Russia UK Ixodes acuminatus France Ixodes canisuga France Germany Spain Ixodes frontalis France Spain Dermacentor reticulatus Germany France UK Germany Haemaphysalis punctata Spain UK Sweden UK Haemaphysalis inermis France Haemaphysalis concinna France Czech Republic Haemaphysalis Portugal marginatum Rhipicephalus sanguineus UK

References

Doby et al. (1990) Postic et al. (1997) Hubbard et al. (1998) Doby et al. (1990) Doby et al. (1991) Liebisch et al. (1998b) Estrada-Peña et al. (1995) Doby et al. (1995) Estrada-Peña et al. (1995) Kahl et al. (1992) Doby et al. (1994) Hubbard et al. (1998) Liebisch et al. (1998b) Marquez and Constan (1990) Nuttall et al. (1994) Tälleklint (1996) Hubbard et al. (1998) Macaigne and Perez-Eid (1991) Doby et al. (1994) Hubálek et al. (1998a) de Michelis et al. (2000) Hubbard et al. (1998)

Argasidae Argas reflexus Argas vespertilionis

Italy UK

Stanek and Simeoni (1989) Hubbard et al. (1998)

INSECTA Siphonaptera (fleas) Ctenophtalmus baeticus arvernus Ctenophtalmus agyrtes Ctenophtalmus solutus

France Czech Republic Slovakia

Doby et al. (1990) Hubálek et al. (1998a) cited in Hubálek et al. (1998a) Doby et al. (1990) cited in Hubálek et al. (1998a) Doby et al. (1991)

Megabothris turbidus

France Slovakia

Spilopsyllus cuniculi

France

Anoplura (sucking lice) Haematopinus suis France

Doby et al. (1994)

Ecology of B. burgdorferi in Europe

157

Table 6.1. Continued Species Diptera Aedes vexans

Aedes cantans Aedes sticticus Culex pipiens pipiens Culex pipiens molestus

Lipoptena cervi

Countries

References

Czech Republic Czech Republic Czech Republic Czech Republic Czech Republic Czech Republic Czech Republic Czech Republic Czech Republic Czech Republic Czech Republic Czech Republic Czech Republic France

Halouzka (1993) Halouzka et al. (1998) Hubálek et al. (1998a) Halouzka et al. (1998) Hubálek et al. (1998a) Halouzka et al. (1998) Hubálek et al. (1998a) Halouzka et al. (1998) Hubálek et al. (1998a) Halouzka (1993) Halouzka et al. (1998) Halouzka et al. (1999) Hubálek et al. (1998a) Doby et al. (1994)

1991). Consequently, each of these hosts could potentially be a reservoir of B. burgdorferi s.l. and therefore the reservoir status of each host species must be individually evaluated. Of about 300 vertebrate species that are hosts for I. ricinus, only a small number have been examined so far for their borrelial infectivity to ticks. Small mammals are certainly the vertebrate group that has been the most extensively investigated up to now, mainly because they can be easily captured, handled and maintained in the laboratory. Several species of mice, voles, rats and shrews have been shown to be competent reservoirs of B. burgdorferi s.l. in Europe (Gern et al., 1998). In particular, strong evidence for reservoir competence has been obtained in many European countries for the mice, Apodemus flavicollis, Apodemus sylvaticus, Apodemus agrarius and for the vole, Clethrionomys glareolus. These rodents transmit B. burgdorferi s.l. to a large number of I. ricinus larvae (Aeschlimann et al., 1986; Matuschka et al., 1992; de Boer et al., 1993; Humair et al., 1993a, 1999; Gern et al., 1994a; Kurtenbach et al., 1994a, 1995, 1998b; Tälleklint and Jaenson, 1994; Randolph and Craine, 1995; Hu et al., 1997; Richter et al., 1999). More limited information has been obtained for other small- and mediumsized mammals. An additional species of vole, Microtus agrestis, has been described as a reservoir in Sweden by Tälleklint and Jaenson (1994), and black rats (Rattus rattus) and Norway rats (Rattus norvegicus) contribute to the transmission of B. burgdorferi s.l. in urbanized environments in continental Europe (Matuschka et al., 1996, 1997). In an urban park in Germany, for example, R. norvegicus rats were less abundant than A. flavicollis, but showed a higher infestation by I. ricinus subadults and a greater infectivity for ticks (Matuschka et al., 1996). Black rats and Norway rats also perpetuate LB spirochaetes in Madeira, a subtropical island where native terrestrial mammals are absent (Matuschka et al., 1994b).

158

L. Gern and P.-F. Humair

Edible dormice (Glis glis) (Matuschka et al., 1994a) and garden dormice (Eliomys quercinus) (Matuschka et al., 1999) have been shown to be reservoir hosts in Germany and in France, respectively. Edible and garden dormice were more heavily infested by larval and nymphal I. ricinus than were mice and voles in the studied habitats. This suggests that in certain habitats where dormice are abundant, these rodent species could be the primary reservoirs (Matuschka et al., 1994b, 1999). Other rodents like grey squirrels (Sciurus carolinensis) in England (Craine et al., 1997) and red squirrels (Sciurus vulgaris) in Switzerland (Humair and Gern, 1998) also contribute to the amplification of the infection in vector ticks. Grey squirrels, which were introduced into Great Britain from North America, were shown to be more important hosts than small rodents for I. ricinus larvae and nymphs (Craine et al., 1995). Moreover, the abundance of grey squirrels makes this species an important source of infection for ticks wherever it is present. Similarly, in continental Europe red squirrels have also been found to be heavily infested with ticks (Matuschka et al., 1996; Humair and Gern, 1998) and feeding ticks showed a high prevalence of infection (69%) (Humair and Gern, 1998). Although red squirrel density has not been evaluated, it is evident that they are potent reservoir hosts and that they contribute to the maintenance of B. burgdorferi s.l. in enzootic areas (Humair and Gern, 1998). Among insectivores, the reservoir competence of three species of shrews (Neomys fodiens, Sorex minutus and Sorex araneus) has been demonstrated in Sweden (Tälleklint and Jaenson, 1994). Evidence of reservoir competence of hedgehogs has been obtained from Ireland (Gray et al., 1994), Germany (Liebisch et al., 1996) and Switzerland (Gern et al., 1997). In Switzerland, an enzootic transmission cycle has been described in an urban environment involving hedgehogs and I. hexagonus in the absence of I. ricinus (Gern et al., 1997). Lagomorphs such as the brown hare (Lepus europaeus) and the varying hare (Lepus timidus) also contribute to the maintenance of B. burgdorferi s.l. in nature, as observed in Sweden in habitats where hares coexist with small mammals (Tälleklint and Jaenson, 1993, 1994), and on islands where hares are the only terrestrial mammal species permanently present (Jaenson and Tälleklint, 1996). Although the European rabbit (Oryctolagus cuniculus) can apparently function as a reservoir host for B. afzelii (Matuschka et al., 2000), the reservoir capacity of this species seems limited since only one out of seven rabbits (14%) was infective to ticks. Among large-sized mammals, two studies in Germany (Kahl and Geue, 1998; Liebisch et al., 1998b) have implicated red foxes as reservoirs, but these animals did not appear to be very potent reservoirs since the borreliae were apparently poorly transmitted to ticks. Birds in Europe were initially identified as non-reservoirs (Matuschka and Spielman, 1992). They were described not only as incompetent reservoirs, but also as having a zooprophylactic role. Zooprophylactic hosts are considered as dead-end hosts for B. burgdorferi s.l. because they eliminate borrelial infection in feeding ticks, which interrupts the maintenance cycle of the spirochaete. Although the involvement of birds has remained controversial for a long time,

Ecology of B. burgdorferi in Europe

159

their role as reservoir hosts is now clearly established. Infected ticks have been collected from various bird species in Sweden (Olsen et al., 1993, 1995a,b), Switzerland (Humair et al., 1993b, 1998), England (Craine et al., 1997; Kurtenbach et al., 1998b) and the Czech Republic (Hubálek et al., 1996). Groundforaging birds such as thrushes (Turdus spp.), blackbirds (Turdus merula), robins (Erithacus rubecula) and pheasants (Phasianus colchicus) were particularly involved, as well as colonial seabirds. B. burgdorferi s.l. was isolated from I. uriae ticks fed on seabirds (Olsen et al., 1993, 1995a; Gylfe et al., 1999), from I. ricinus ticks fed on blackbirds (Hubálek et al., 1996; Humair et al., 1998), from the skin of blackbirds (T. merula) (Humair et al., 1998) and from the blood of puffins (Fratercula artica) (Gylfe et al., 1999), suggesting that live spirochaetes can survive in birds and in bird-feeding ticks. Strong evidence of reservoir competence was eventually obtained for two bird species: blackbird (T. merula) (Humair et al., 1998) and pheasant (P. colchicus) (Kurtenbach et al., 1998a) were investigated by tick xenodiagnosis, which clearly demonstrated the efficient transmission of borrelial infection from birds to ticks. B. burgdorferi s.l. also circulates between seabirds such as razorbills (Alca torda) and puffins (F. artica) and the seabird tick I. uriae in isolated environments (Olsen et al., 1993, 1995a; Gylfe et al., 1999). Borrelia-infected I. uriae ticks have been observed in seabird colonies in the northern and southern hemispheres (Olsen et al., 1995a). The similarity of DNA sequences of B. burgdorferi s.l. from I. uriae of various locations strongly suggests that seabirds transfer borrelial spirochaetes between seabird colonies (Olsen et al., 1995a). Reactivation of latent borrelial infection has been observed with passerines experimentally submitted to stressful conditions that simulate migration (progressive decrease of photoperiod) (Gylfe et al., 2000). This suggests that migrating birds could spread B. burgdorferi s.l. spirochaetes to ticks encountered along the migration route. The determination of the reservoir competence of a host species is important but should also include quantification of the contribution of a particular host species in a particular habitat. The immune status of the host may have an impact on the circulation of B. burgdorferi s.l. in nature by affecting the ability of ticks to acquire infection from reservoir hosts. Apodemus spp. mice and Clethrionomys spp. voles are a good illustration of this phenomenon. These host species seem to have developed different responses towards tick infestation and borrelial infection, and this influences their respective reservoir importance. Kurtenbach et al. (1994b) showed in a laboratory study that C. glareolus voles express a low immunity to spirochaetes and as a consequence develop high levels of infection and show high transmission rates to ticks. In contrast, Apodemus spp. mice control spirochaete infection more effectively by a specific immune response that maintains their borrelial infection at a low level and also at a low transmission rate to ticks (Kurtenbach et al., 1994b). In parallel, Dizij and Kurtenbach (1995) demonstrated experimentally that voles, but not mice, acquire resistance to I. ricinus ticks with consecutive infestations. Interestingly, some of the phenomena demonstrated by these laboratory results have been confirmed under natural conditions. Spirochaetes are more easily isolated from ear biopsies of Clethrionomys

160

L. Gern and P.-F. Humair

spp. voles than of Apodemus spp. mice (Petney et al., 1996; Humair et al., 1999). This high level of B. burgdorferi infection in voles also allows a higher transmission rate of spirochaetes from voles to the ticks feeding on them (Humair et al., 1999). However, voles progressively acquire resistance to ticks after several tick infestations. Ticks that feed on these resistant voles suffer a retarded development due to a reduction of their engorgement weights as well as an increased mortality since they do not moult successfully. The low moulting success of ticks fed on voles lessens the relative contribution of voles in endemic areas, despite their high infection rate and their high infectivity for ticks (Humair et al., 1999). Reservoir potential (as defined by Mather et al., 1989) or coefficient of transmission (Randolph and Craine, 1995) is an attempt to quantify the contribution as a reservoir of a particular host species in a particular habit. The reservoir potential is influenced by three main factors: (i) host infectivity to ticks; (ii) degree of tick–host contact (mean infestation rate by larval ticks); and (iii) host density in the habitat (Mather et al., 1989). The higher the values of these three parameters, the higher the reservoir potential of the host and the more important its contribution as a reservoir in the studied habitat. The evaluation of reservoir potential is not an easy task since it requires the assessment of various factors (Tälleklint and Jaenson, 1994), but this multifaceted approach should be encouraged in the future. The determination of a vertebrate species as a reservoir is obviously important in understanding the ecology of LB. The identification of a host as a non-reservoir or as a zooprophylactic host is also important since the circulation of B. burgdorferi s.l. in nature is impaired by such hosts. The identification of a nonreservoir is in fact much more difficult than that of a reservoir host, since repeated investigations are required. Until now, only a few vertebrate species have been described as non-reservoirs or zooprophylactic hosts in Europe. The lizard (Lacerta viridis) was reported to have a zooprophylactic effect on the island of Madeira since spirochaetes were destroyed in the midgut of ticks feeding on this host (Matuschka et al., 1994b). Several other vertebrates such as blackbirds (Matuschka and Spielman, 1992) and some ungulates (roe deer, C. capreolus; red deer, Cervus elaphus; sika deer, Cervus nippon (cited by Gern et al., 1998); moose, Alces alces; fallow deer, Dama dama; cattle, Bos taurus; sheep, Ovis aries) (Gray et al., 1992, 1995; Jaenson and Tälleklint, 1992; Tälleklint and Jaenson, 1994) have been reported as zooprophylactic or non-reservoir hosts. However, further studies have shown that some of these host species are in fact reservoir-competent for some genospecies of B. burgdorferi, for example blackbirds (Humair et al., 1998), and others, for example sheep (Ogden et al., 1997) and sika deer (Kimura et al., 1995), may also contribute to the circulation of B. burgdorferi s.l. in nature. Various factors may lead to the conclusion that a host is non-reservoir or is zooprophylactic. For example, most methods for the determination of the reservoir competence of a host are based on detection of systemic infections. However, co-feeding transmission, in which uninfected engorging ticks may acquire borrelia infections from localized sites where infected ticks feed simultaneously (Gern and Rais, 1996; Randolph et al., 1996), has been reported in sheep

Ecology of B. burgdorferi in Europe

161

in Britain (Ogden et al., 1997) and sika deer in Japan (Kimura et al., 1995). Even if some ungulates do not sustain systemic borrelia infection, those in which cofeeding transmission occurs can be considered as amplifying hosts. Quantification of co-feeding transmission in ungulates, particularly in deer, and its importance in maintaining disease foci remain to be evaluated. Another factor complicating the determination of reservoir competence of a vertebrate host is the possibility of host specificity of different Borrelia genospecies. This aspect has been neglected for a long time and even now knowledge of the relationships between the different B. burgdorferi s.l. species and their hosts is poorly documented. Indications of this phenomenon in Europe were first provided by the research of Humair et al. (1995) in Switzerland, who showed that borrelial isolates from 14 ear biopsies taken from ten C. glareolus voles and one A. flavicollis mouse (three isolates were obtained from two recaptured voles) captured in two endemic areas belonged exclusively to B. afzelii, although B. burgdorferi s.s. and B. garinii were present in questing ticks in the same areas. Interestingly, the first borrelial isolate obtained from a rodent in Sweden (Hovmark et al., 1988) was characterized as B. afzelii (Postic et al., 1994). More recently, B. afzelii was once again isolated from ear biopsies of mice and voles collected in a third study site in Switzerland, suggesting that a specific association occurs between B. afzelii and small rodents (Humair et al., 1999). The use of tick xenodiagnosis on mice and voles showed that B. afzelii is clearly the dominant genospecies transmitted from small rodents to I. ricinus ticks in two enzootic areas in Switzerland (Hu et al., 1997; Humair et al., 1999). Out of 124 positive xenodiagnostic ticks, 120 ticks (97%) were infected by B. afzelii, two ticks contained B burgdorferi s.s. (one with a co-infection with B. afzelii) and two ticks contained uncharacterized spirochaetes. The isolation of B. afzelii from xenodiagnostic ticks demonstrates the viability of spirochaetes discovered in rodent-feeding ticks, which is a crucial point in the transmission of B. burgdorferi s.l. from rodents. All these results strongly suggest that an association exists between Apodemus spp. mice, Clethrionomys spp. voles and B. afzelii, since this genospecies is almost exclusively transmitted to I. ricinus larvae from these animals. Dolan et al. (1998) obtained similar results with I. ricinus ticks and laboratory mice in a Centres for Disease Control and Prevention (CDC) laboratory, where B. afzelii was isolated in all ear biopsy cultures from outbred laboratory mice that had been infested with I. ricinus ticks collected from the field near Neuchâtel (Switzerland). Additionally, xenodiagnostic I. ricinus larvae fed on outbred mice inoculated with other strains of B. burgdorferi s.l. (B. afzelii strain PGau.C3, B. garinii strains VS286 and VSBP, B. burgdorferi s.s. strains B-31 and B-31.D1) acquired all three Borrelia genospecies. However, the prevalence of B. afzelii infection in ticks was particularly high (90%) compared with low infection prevalences (3–5%) with B. garinii and B. burgdorferi s.s. Further, only B. afzelii survived the transstadial maintenance to nymphs (Dolan et al., 1998). Additional observations were obtained by Hubálek et al. (1998a), who isolated B. afzelii from two A. flavicollis, one C. glareolus and two fleas (Ctenophthalmus agyrtes) fed on C. glareolus in the Czech Republic. In parallel with the observations made by Humair et al. (1995) on B. afzelii and

162

L. Gern and P.-F. Humair

small rodents, Olsen et al. (1995b) described the predominant presence of B. garinii in ticks feeding on migrating birds collected in Sweden and Denmark: B. garinii DNA was found in 70% of I. ricinus larvae (23/33), B. burgdorferi s.s. DNA in 20% (7/33) and B. afzelii DNA in 9% (3/33). B. garinii was detected in all I. ricinus larvae (n = 11) collected from autumn-migrating birds and was also most prevalent in I. ricinus larvae collected from spring-migrating birds. In addition to migrating passerines, the existence of an association between birds and B. garinii was clearly illustrated by Olsen et al. (1995a) in the case of seabirds and the vector I. uriae. In the isolated maintenance cycle involving seabirds and the seabird tick (Olsen et al., 1993), B. garinii was isolated and characterized from I. uriae ticks and more recently from seabird tissues (Olsen et al., 1995a; Gylfe et al., 1999). In pheasants, more than 50% of I. ricinus ticks were infected by borreliae, particularly B. garinii (27%) and B. valaisiana (16%) (Kurtenbach et al., 1998b). Hubálek et al. (1996) in the Czech Republic obtained a B. garinii isolate from a nymphal I. ricinus fed on T. merula. Shortly afterwards, B. garinii and B. valaisiana infections (seven isolates and four DNA detections) were revealed in five T. merula and one Turdus philomelos (Humair et al., 1998). No infection with Borrelia genospecies other than B. garinii and B. valaisiana has been observed in European Turdus spp. Characterization of borreliae from xenodiagnostic ticks fed on blackbirds in Switzerland demonstrated that B. valaisiana was most prevalent in these ticks (30%) when compared with B. garinii (3%) and B. afzelii (3%). Co-infections with B. valaisiana and B. garinii or B. afzelii were observed in 4% of the ticks. Interestingly, only B. valaisiana and B. garinii were isolated successfully from these xenodiagnostic ticks, whereas the presence of B. afzelii was revealed by DNA detection only (Humair et al., 1998). This shows that blackbirds transmitted viable B. valaisiana and B. garinii to ticks feeding on them whereas the viability of B. afzelii spirochaetes in bird-feeding ticks remains uncertain. More recently, Gylfe et al. (2000) demonstrated that B. garinii latent infection can be reactivated in experimentally infected redwing thrushes (Turdus iliacus) placed under stressful conditions simulating migration and causing migratory restlessness. This suggests that Turdus spp. passerines can sustain B. garinii infection for a long time and that migrating birds placed under conditions of autumnal migration transmit B. garinii to ticks. This corroborates the previous observation of B. garinii in all examined I. ricinus ticks from autumnal migrant birds (Olsen et al., 1995b). Birds have a high body temperature, and Hubálek et al. (1998b) reported that the optimal growth temperature of B. garinii is higher than that of the other two pathogenic species, B. afzelii and B. burgdorferi s.s. Overall, these data strongly suggest that an association exists between birds and B. garinii and also B. valaisiana, a more recently described genospecies. Interestingly, an association between B. garinii and birds has been observed in Japan (Nakao et al., 1994; Miyamoto et al., 1997; Ishiguro et al., 2000). However, in areas outside the geographical distribution of B. garinii, such as North America, B. burgdorferi s.s. is the Borrelia sp. circulating between birds and ticks (Richter et al., 2000).

Ecology of B. burgdorferi in Europe

163

In addition to the associations of small rodents with B. afzelii, and birds with B. garinii and B. valaisiana, another association has been described in Switzerland and in the UK. As previously mentioned, grey and red squirrels (S. carolinensis and S. vulgaris) are reservoirs for B. burgdorferi s.l. In Switzerland, B. burgdorferi s.s. and B. afzelii were the only two genospecies isolated from skin samples of red squirrels (S. vulgaris) and were the two most prevalent genospecies found in I. ricinus ticks engorging on red squirrels, representing 96% of the isolates (Humair and Gern, 1998). The high frequency of B. burgdorferi s.s. and B. afzelii in the skin of squirrels as well as in squirrel-feeding I. ricinus suggests that both genospecies are preferentially transmitted from squirrels to ticks. Similar findings were obtained by Craine et al. (1997) in the UK: B. afzelii was identified in I. ricinus nymphs fed on a naturally infected grey squirrel and another grey squirrel sustained an experimental infection with B. burgdorferi s.s. However, other borrelia–host associations that may depend on the local ecosystems have been described. In some parts of Europe, in addition to B. afzelii, B. garinii was also observed to be associated with rodents. Thus, data from Austria, Germany and Russia showed that different genospecies can be isolated from or DNA detected (by PCR) in the urinary bladder, heart and spleen of small mammals or in xenodiagnostic ticks fed on rats and yellow-necked mice (Khanakah et al., 1994; Gorelova et al., 1995; Richter et al., 1999). Differences in biological and ecological factors occurring in some parts of Europe, such as the presence of different tick and rodent species or of different B. garinii subtypes, could explain the observed discrepancy. For example, B. garinii serotype 4, which has only been described in the cerebrospinal fluid of patients in Germany, The Netherlands, Denmark and Slovenia (Wilske et al., 1993, 1996; van Dam et al., 1997), is apparently rare in nature and has so far only been found in I. ricinus ticks collected in Germany (Eiffert et al., 1995; Hu et al., 2001). The difficulty in obtaining infected ticks from laboratory mice infected with B. garinii (usually associated with birds) has been documented previously (Gern et al., 1994b; Dolan et al., 1998). However, it was shown recently that laboratory mice infected with the serotype 4 of B. garinii easily transmitted this serotype to ticks (Hu et al., 2001), which could be explained by the resistance of this serotype to complement (van Dam et al., 1997). Therefore, one explanation for the observation of B. garinii in rodents in Austria, Germany and Russia may be the presence of B. garinii serotype 4 or related serotypes in these areas. In Japan, a similar association has been described between rodents and B. garinii ribotype IV, since this particular ribotype was predominant in rodents (Apodemus speciosus and Clethrionomys rutilus) and rodent-feeding I. persulcatus ticks (Nakao et al., 1994). These observations show how diverse the ecology of B. burgdorferi s.l. can be. Although some of the reservoirs of B. afzelii, B. burgdorferi s.s., B. garinii and B. valaisiana have been documented, those of B. lusitaniae still remain to be identified. De Michelis et al. (2000) suggested that avian and rodent species might be reservoir hosts for B. lusitaniae because DNA of this genospecies has been detected in adults of Hyalomma marginatum, a tick species that mainly feeds on birds and rodents as subadults (Hillyard, 1996). Other borrelia types identified

164

L. Gern and P.-F. Humair

in European patients by Picken et al. (1996) and Wang et al. (1999) have not been described in either ticks or wild animals. Specific associations between reservoirs and Borrelia spp. spirochaetes imply that in certain cases B. burgdorferi s.l. genospecies are also transmitted to inadequate hosts. The specific association between birds and B. garinii/B. valaisiana in Europe may explain why the reservoir role of birds has remained controversial for a long time and why they were first described as zooprophylactic hosts. The first European studies on birds were performed before the description of different genospecies within the B. burgdorferi s.l. species complex. The existence of specific associations explains why blackbirds (T. merula) were first described as incompetent reservoirs and even as zooprophylactic hosts for LB spirochaetes by Matuschka and Spielman (1992). In this experiment, blackbirds were exposed to infected I. ricinus nymphs that acquired infection after feeding as larvae on black-striped mice (A. agrarius). Thus, considering the association of small rodents and B. afzelii, it is most probable that the I. ricinus nymphs used to infect blackbirds carried mouse-adapted Borrelia genospecies. Consequently, blackbirds were effectively inadequate hosts for the Borrelia spp. to which they had been exposed, because they are incompetent reservoirs and zooprophylactic hosts for rodentadapted genospecies. The existence of specific spirochaetal transmission patterns, of zooprophylactic and of non-reservoir hosts can be explained by the specific borreliacidal effect of host complement (Kurtenbach et al., 1998c). This effect may prevent infection of the vertebrate host by some Borrelia species and may result in loss of spirochaetes by ticks that feed on inadequate hosts. The pattern of serum sensitivity of different Borrelia genospecies matches the known reservoir status of many vertebrate species for B. burgdorferi s.l. For example, B. afzelii was resistant to rodent sera whereas a strain of B. garinii was readily lysed, and the absence of systemic infection in deer correlates with the borreliacidal activity of deer sera against all the genospecies tested (Kurtenbach et al., 1998c).

Conclusions The ecological picture of LB is particularly complex because of the existence of diversity within the three main components of this zoonosis: pathogens, vectors and reservoirs. At present in Europe, five B. burgdorferi s.l. genospecies, three vector species and about 35 reservoir host species are linked in a web of relationships between each other. By focusing on this web, it can be seen that the links are organized, since each element of the web has privileged relationships with some others. For example, the seabird tick, I. uriae, is only infected by B. garinii, whereas I. ricinus may be infected by any of the five genospecies present in Europe. Similarly, I. uriae exclusively infests seabirds, whereas I. ricinus infests a large range of mammals and birds. It is obvious from this example that B. garinii is associated with seabirds, whereas similar associations between hosts of I. ricinus and B. burgdorferi s.l. genospecies are less obvious.

Ecology of B. burgdorferi in Europe

165

Although I. ricinus infection by B. burgdorferi s.l. genospecies and host infestation by ticks are generally well documented, the relationships between I. ricinus hosts and borreliae needs clarification in many respects. Additionally, the web varies geographically as well as temporally. A huge amount of knowledge has been collected on the ecology of B. burgdorferi s.l. in Europe but further studies are needed for a better understanding of the situation. Emphasis has to be put on the distribution of the various Borrelia species and subtypes throughout Europe, the identification of additional reservoir hosts, their relationships with the various Borrelia species and the immune response of hosts towards ticks and borreliae. Because of regional and local pecularities, these studies should be performed in as many habitats as possible. The collection of these local snapshots will help towards a better picture of the ecology of B. burgdorferi s.l. in Europe.

References Aeschlimann, A. (1972) Ixodes ricinus, Linné, 1758 (Ixodoidea: Ixodidae). Essai préliminaire de synthèse sur la biologie de cette espèce en Suisse. Acta Tropica 29, 321–340. Aeschlimann, A., Chamot, E., Gigon, F., Jeanneret, J.P., Kesseler, D. and Walther, C. (1986) B. burgdorferi in Switzerland. Zentralblatt für Bakteriologie, Mikrobiologie und Hygiene A 263, 450–458. Anderson, J.F. (1991) Epizootiology of Lyme borreliosis. Scandinavian Journal of Infectious Diseases Suppl. 77, 23–34. Arthur D.R. (1953) The host relationships of Ixodes hexagonus Leach in Britain. Parasitology 43, 227–238. Arthur, D.R. (1968) British Ticks. Butterworths, London, 213 pp. Baranton, G., Postic, D., Saint Girons, I., Boerlin, P., Piffaretti, J.C., Assous, M. and Grimont, P.A.D. (1992) Delineation of Borrelia burgdorferi sensu stricto, Borrelia garinii sp. nov., and group VS461 associated with Lyme borreliosis. International Journal of Systematic Bacteriology 42, 378–383. Bellet-Edimo, O.R. (1997) Importance de la transmission transstadiale et de la transmission transovarienne du spirochète Borrelia burgdorferi (Spiro-

chaetales: Spirochaetaceae) chez la tique Ixodes ricinus (Acari: Ixodidae) dans l’épidémiologie de la borréliose de Lyme. PhD thesis, University of Neuchâtel, Neuchâtel, Switzerland. Bernasconi, M.V., Valsangiacomo, C., Balmelli, T., Péter, O. and Piffaretti, J.C. (1997) Tick zoonoses in the southern part of Switzerland (Canton Ticino): occurrence of Borrelia burgdorferi sensu lato and Rickettsia sp. European Journal of Epidemiology 13, 209–215. de Boer, R., Hovius, K.E., Nohlmans, M.K.E. and Gray, J.S. (1993) The woodmouse (Apodemus sylvaticus) as a reservoir of tick-transmitted spirochetes (Borrelia burgdorferi) in the Netherlands. Zentralblatt für Bakteriologie 279, 404–416. Burgdorfer, W., Barbour, A.G., Hayes, S.F., Benach, J.L., Grunwaldt, E. and Davies, J.P. (1982) Lyme disease – a tick-borne spirochetosis? Science 220, 321–322. Burgdorfer, W., Barbour, A.G., Hayes, S.F., Péter, O. and Aeschlimann, A. (1983) Erythema chronicum migrans – a tick-borne spirochetosis. Acta Tropica 40, 79–83. Canica, M.M., Nato, F., du Merle, L., Mazie, J.C., Baranton, G. and Postic,

166

L. Gern and P.-F. Humair

D. (1993) Monoclonal antibodies for identification of Borrelia afzelii sp. nov. associated with late cutaneous manifestations of Lyme borreliosis. Scandinavian Journal of Infectious Diseases 25, 441–448. Cinco, M., Padovan, D., Murgia, R., Poldini, L., Frusteri, L., van de Pol, I., Verbeek-De Kruif, N., Rijpkema, S.G. and Marioli, M. (1998) Rate of infection of Ixodes ricinus ticks with Borrelia burgdorferi sensu stricto, Borrelia garinii, Borrelia afzelii and group VS116 in an endemic focus of Lyme disease in Italy. European Journal of Clinical Microbiology and Infectious Diseases 17, 90–94. Craine, N.G., Randolph, S.E. and Nuttall P.A. (1995) Seasonal variation in the role of grey squirrels as hosts of Ixodes ricinus, the tick vector of the Lyme disease spirochaete, in a British woodland. Folia Parasitologica 42, 73–80. Craine, N.G., Nuttall, P.A., Marriott, A.C. and Randolph, S.E. (1997) Role of grey squirrels and pheasants in the transmission of Borrelia burgdorferi sensu lato, the Lyme disease spirochaete, in the UK. Folia Parasitologica 44, 155–160. Cutler, S.J., Williams, N.G. and Wright, D.J.M. (1989) Borrelia species from an endemic area for Lyme disease in the United Kingdom. In: 4th European Congress of Clinical Microbiology, Nice, France, Abstract 908. van Dam, A.P., Oei, A., Jaspars, R., Fijen, C., Wilske, B., Spanjaard, L. and Dankert, J. (1997) Complement-mediated serum sensitivity among spirochetes that cause Lyme disease. Infection and Immunity 65, 1228–1236. Dizij, A. and Kurtenbach, K. (1995) Clethrionomys glareolus, but not Apodemus flavicollis, acquires resistance to Ixodes ricinus L., the main European vector of Borrelia burgdorferi. Parasite Immunology 17, 177–183. Doby, J.M., Bigaignon, G., Launay, H., Costil, C. and Lorvellec, O. (1990) Présence de Borrelia burgdorferi, agent de spirochétoses à tiques, chez Ixodes

(Exopalpiger) trianguliceps Birula, 1895 et Ixodes (Ixodes) acuminatus Neumann, 1901 (Acariens Ixodidae) et chez Ctenophtalmus baeticus arvernus Jordan, 1931 et Megabothris turbidus Rothschild, 1909 (Insectes Siphonaptera), ectoparasites de micromammifères des forêts dans l’Ouest de la France. Bulletin de la Société Française de Parasitologie 8, 311–322. Doby, J.M., Bigaignon, G., Aubert, M. and Imbert, G. (1991) Ectoparasites du renard et borréliose de Lyme. Recherche de Borrelia burgdorferi chez les tiques Ixodidae et insectes Siphonaptera. Bulletin de la Société Française de Parasitologie 9, 279–288. Doby, J.M., Bigaignon, G., Degeilh, B. and Guigen, C. (1994) Ectoparasites des grands mammifères sauvages, cervidés et suidés, et borréliose de Lyme. Recherche de Borrelia burgdorferi chez plus de 1400 tiques, poux, pupipares et puces. Revue Médicale et Vétérinaire 145, 743–748. Doby, J.M., Degeilh, B., Cayouette, S. and Guigen, C. (1995) Présence de Borrelia burgdorferi sensu lato chez Ixodes (Trichotoixodes) pari Leach, 1815 (Acari: Ixodidae), tique strictement spécifique des oiseaux. Bulletin de la Société de Pathologie Exotique 88, 185–186. Dolan, M.C., Piesman, J., Mbow, M.L., Maupin, G.O., Péter, O., Brossard, M. and Golde, W.T. (1998) Vector competence of Ixodes scapularis and Ixodes ricinus (Acari: Ixodidae) for three genospecies of Borrelia burgdorferi. Journal of Medical Entomology 35, 465–470. Eiffert, H., Ohlenbusch, A., Christen, H.J., Thomssen, R., Spielman, A. and Matuschka, F.R. (1995) Nondifferentiation between Lyme disease spirochetes from vector ticks and human cerebrospinal fluid. Journal of Infectious Diseases 171, 476–479. Eley, T.J. (1977) Ixodes uriae (Acari, Ixodidae) from a river otter. Journal of Medical Entomology 13, 506. Escudero, R., Barral, M., Pérez, A.,

Ecology of B. burgdorferi in Europe

Vitutia, M.M., Garcia-Perez, A.L., Jimenez, S., Sellek, R.E. and Anda, P. (2000) Molecular and pathogenic characterization of Borrelia burgdorferi sensu lato isolates from Spain. Journal of Clinical Microbiology 38, 4026–4033. Estrada-Peña, A., Oteo, J.A., EstradaPeña, R., Gortázar, C., Osácar, J.J., Moreno, J.A. and Castellá, J. (1995) Borrelia burgdorferi sensu lato in ticks (Acari: Ixodidae) from two different foci in Spain. Experimental and Applied Acarology 19, 173–180. Fukunaga, M., Hamase, A., Okada, K. and Nakao, M. (1996) Borrelia tanuki sp. nov. and Borrelia turdae sp. nov. found from ixodid ticks in Japan: rapid species identification by 16S rRNA. Microbiology and Immunology 40, 877–881. Gern, L. and Rais, O. (1996) Efficient transmission of Borrelia burgdorferi between cofeeding Ixodes ricinus ticks (Acari: Ixodidae). Journal of Medical Entomology 33, 189–192. Gern, L., Toutoungi, L.N., Hu, C.M. and Aeschlimann, A. (1991) Ixodes (Pholeoixodes) hexagonus, an efficient vector of Borrelia burgdorferi in the laboratory. Medical and Veterinary Entomology 5, 431–435. Gern, L., Siegenthaler, M., Hu, C.M., Leuba-Garcia, S., Humair, P.F. and Moret, J. (1994a) Borrelia burgdorferi in rodents (Apodemus flavicollis and A. sylvaticus): duration and enhancement of infectivity for Ixodes ricinus ticks. European Journal of Epidemiology 10, 75–80. Gern, L., Rais, O., Capiau, C., Hauser, P., Lobet, Y., Simoen, E., Voet, P. and Petre, J. (1994b) Immunization of mice by recombinant OspA preparations and protection against Borrelia burgdorferi infection induced by Ixodes ricinus tick bites. Immunology Letters 39, 249–258. Gern, L., Rouvinez, E., Toutoungi, L.N. and Godfroid, E. (1997) Transmission cycles of Borrelia burgdorferi sensu lato involving Ixodes ricinus and/or I. hexagonus ticks and the European hedgehog, Erinaceus europaeus, in suburban and

167

urban areas in Switzerland. Folia Parasitologica 44, 309–314. Gern, L., Estrada-Peña, A., Frandsen, F., Gray, J.S., Jaenson, T.G.T., Jongejan, F., Kahl, O., Korenberg, E., Mehl, R. and Nuttall, P.A. (1998) European reservoir hosts of Borrelia burgdorferi sensu lato. Zentralblatt für Bakteriologie 287, 196–204. Gern, L., Hu, C.M., Kocianova, E.,Vyrostekova, V. and Rehacek, J. (1999) Genetic diversity of Borrelia burgdorferi sensu lato isolates obtained from Ixodes ricinus ticks collected in Slovakia. European Journal of Epidemiology 15, 665–669. Gigon, F. (1985) Biologie d’Ixodes ricinus L. sur le Plateau Suisse – une contribution à l’écologie de ce vecteur. PhD thesis, University of Neuchâtel, Neuchâtel, Switzerland. Gorelova, N.B., Korenberg, E.I., Kovalevskii, Y.V. and Shcherbakov, S.V. (1995) Small mammals as reservoir hosts for borrelia in Russia. Zentralblatt für Bakteriologie 282, 315–322. Gray, J.S. (1991) The development and seasonal activity of the tick Ixodes ricinus: a vector of Lyme borreliosis. Review of Medical and Veterinary Entomology 79, 323–333. Gray, J.S., Kahl, O., Janetzki, C. and Stein, J. (1992) Studies on the ecology of Lyme disease in a deer forest in County Galway, Ireland. Journal of Medical Entomology 29, 915–920. Gray, J.S., Kahl, O., Janetzki-Mittman, C., Stein, J. and Guy, E. (1994) Acquisition of Borrelia burgdorferi by Ixodes ricinus ticks fed on the European hedgehog, Erinaceus europaeus L. Experimental and Applied Acarology 18, 485–491. Gray, J.S., Kahl, O., Janetzki, C., Stein, J. and Guy, E. (1995) The spatial distribution of Borrelia burgdorferi-infected Ixodes ricinus in the Connemara region of Co. Galway, Ireland. Experimental and Applied Acarology 18, 485–491. Gray, J.S., Kahl, O., Robertson, J.N.,

168

L. Gern and P.-F. Humair

Daniel, M., Estrada-Pena, A., Gettinby, G., Jaenson, T.G.T., Jaenson, P., Jongejan, F., Korenberg, E., Kurtenbach, K. and Zeman, P. (1998) Lyme borreliosis habitat assessment. Zentralblatt für Bakteriologie und Hygiene 287, 211–228. Gylfe, Å., Olsen, B., Strasevicius, D., Ras, N.M., Weihe, P., Noppa, L., Ostberg, Y., Baranton, G. and Bergström, S. (1999) Isolation of Lyme disease Borrelia from puffins (Fratercula arctica) and seabird ticks (Ixodes uriae) on the Faeroe Islands. Journal of Clinical Microbiology 37, 890–896. Gylfe, Å., Bergström, S., Lunström, J. and Olsen, B. (2000) Reactivation of Borrelia infection in birds. Nature 403, 724–725. Halouzka, J. (1993) Borreliae in Aedes vexans and hibernating Culex pipiens molestus mosquitoes. Biológia 48, 123–124. Halouzka, J., Postic, D. and Hubálek, Z. (1998) Isolation of the spirochaete Borrelia afzelii from the mosquito Aedes vexans in the Czech Republic. Medical and Veterinary Entomology 12, 103–105. Halouzka, J., Wilske, B., Stünzer, D., Sanogo, Y.O. and Hubálek, Z. (1999) Isolation of Borrelia afzelii from overwintering Culex pipiens biotype molestus mosquitoes. Infection 27, 275–277. Hillyard, P.D. (1996) Ticks of North-west Europe, Vol. 52. Edited by Barnes, R.S.K. and Crothers, J.H. Field Studies Council, Shrewsbury, UK, 178 pp. Hovmark, A., Jaenson, T.G.T., Åsbrink, E., Forsman, A. and Jansson, E. (1988) First isolations of Borrelia burgdorferi from rodents collected in Northern Europe. Acta Pathologica, Microbiologica et Immunologica Scandinavica Section B 96, 917–920. Hu, C.M., Humair, P.F., Wallich, R. and Gern, L. (1997) Apodemus sp. rodents, reservoir hosts for Borrelia afzelii in an endemic area in Switzerland. Zentralblatt für Bakteriologie 285, 558–564. Hu, C.M., Wilske, B., Fingerle, V., Lobet, Y. and Gern, L. (2001) Transmission of Borrelia garinii OspA serotype 4 to

BALB/c mice by Ixodes ricinus ticks collected in the field. Journal of Clinical Microbiology 39, 1169–1171. Hubálek, Z. and Halouzka, J. (1997) Distribution of Borrelia burgdorferi sensu lato genomic groups in Europe, a review. European Journal of Epidemiology 13, 951–957. Hubálek, Z. and Halouzka, J. (1998) Prevalence rates of Borrelia burgdorferi sensu lato in host-seeking Ixodes ricinus ticks in Europe. Parasitology Research 84, 167–172. Hubálek, Z., Anderson, J.F., Halouzka, J. and Hájek, V. (1996) Borreliae in immature Ixodes ricinus (Acari: Ixodidae) ticks parasitizing birds in the Czech Republic. Journal of Medical Entomology 33, 766–771. Hubálek, Z., Halouzka, J. and Juricová, Z. (1998a) Investigation of haematophagous arthropods for borreliae – summarized data, 1988–1996. Folia Parasitologica 45, 67–72. Hubálek, Z., Halouzka, J. and Heroldova, M. (1998b) Growth temperature ranges of Borrelia burgdorferi sensu lato strains. Journal of Medical Microbiology 47, 929–932. Hubbard, M.J., Baker, A.S. and Cann, K.J. (1998) Distribution of Borrelia burgdorferi s.l. spirochaete DNA in British ticks (Argasidae and Ixodidae) since the 19th century, assessed by PCR. Medical and Veterinary Entomology 12, 89–97. Humair, P.F. and Gern, L. (1998) Relationship between Borrelia burgdorferi sensu lato species, red squirrels (Sciurus vulgaris) and Ixodes ricinus in enzootic areas in Switzerland. Acta Tropica 69, 213–227. Humair, P.F., Turrian, N., Aeschlimann, A. and Gern, L. (1993a) Borrelia burgdorferi in a focus of Lyme borreliosis: epizootiologic contribution of small mammals. Folia Parasitologica 40, 65–70. Humair, P.F., Turrian, N., Aeschlimann, A. and Gern, L. (1993b) Ixodes ricinus immatures on birds in a focus of

Ecology of B. burgdorferi in Europe

Lyme borreliosis. Folia Parasitologica 40, 237–242. Humair, P.F., Péter, O., Wallich, R. and Gern, L. (1995) Strain variation of Lyme disease spirochetes isolated from Ixodes ricinus ticks and rodents collected in two endemic areas in Switzerland. Journal of Medical Entomology 32, 433–438. Humair, P.F., Postic, D., Wallich, R. and Gern, L. (1998) An avian reservoir (Turdus merula) of the Lyme borreliosis spirochetes. Zentralblatt für Bakteriologie 287, 521–538. Humair, P.F., Rais, O. and Gern, L. (1999) Transmission of Borrelia afzelii from Apodemus mice and Clethrionomys voles to Ixodes ricinus ticks: differential transmission pattern and overwintering maintenance. Parasitology 118, 33–42. Ishiguro, F., Takada, N., Masuzawa, T. and Fukui, T. (2000) Prevalence of Lyme disease Borrelia spp. in ticks from migratory birds on the Japanese mainland. Applied and Environmental Microbiology 66, 982–986. Jaenson, T.G.T. and Tälleklint, L. (1992) Incompetence of roe deer as reservoirs of the Lyme borreliosis spirochete. Journal of Medical Entomology 29, 813–817. Jaenson, T.G.T. and Tälleklint, L. (1996) Lyme borreliosis spirochetes in Ixodes ricinus (Acari: Ixodidae) and the varying hare on isolated islands in the Baltic sea. Journal of Medical Entomology 33, 339–343. Jauris-Heipke, S., Liegl, G., Preac-Mursic, V., Rössler, D., Schwab, E., Soutschek, E., Will, G. and Wilske, B. (1995) Molecular analysis of genes encoding outer surface protein C (OspC) of Borrelia burgdorferi sensu lato: relationship to ospA genotype and evidence of lateral gene exchange of ospC. Journal of Clinical Microbiology 33, 1860–1866. Johnson, R.C., Schmid, G.P., Hyde, F.W., Steigerwalt, A.G. and Brenner, D.J. (1984) Borrelia burgdorferi sp. nov.: etio-

169

logic agent of Lyme disease. International Journal of Systematic Bacteriology 34, 496–497. Kahl, O. and Geue, L. (1998) Laboratory study on the possible role of the European fox, Vulpes vulpes as a potential reservoir of Borrelia burgdorferi s.l. In: Coons, L. and Rothschild, M. (eds) 2nd International Conference on Tick-Borne Pathogens at the Host–Vector Interface: a Global Perspective, S.I. The Conference, Kruger National Park, South Africa, p. 29. Kahl, O., Janetzki, C., Gray, J.S., Stein, J. and Bauch, R.J. (1992) Tick infection rates with Borrelia: Ixodes ricinus versus Haemaphysalis concinna and Dermacentor reticulatus in two locations in eastern Germany. Medical and Veterinary Entomology 6, 363–366. Kawabata, H., Masuzawa, T. and Yanagihara, Y. (1993) Genomic analysis of Borrelia japonica sp. nov. isolated from Ixodes ovatus in Japan. Microbiology and Immunology 37, 843–848. Khanakah, G., Kmety, E., Radda, A. and Stanek, G. (1994) Micromammals as reservoir of Borrelia burgdorferi in Austria. In: Cevenini, R., Sambri, V. and La Placa, M. (eds) Abstract Book of the VI Conference on Lyme Borreliosis. Bologna, Italy, Abstract P077W. Kimura, K., Isogai, E., Isogai, H., Kamewaka, Y., Nishikawa, T., Ishii, N. and Fujii, N. (1995) Detection of Lyme disease spirochetes in the skin of naturally infected wild sika deer (Cervus nippon yeosoensis) by PCR. Applied and Environmental Microbiology 61, 1641–1642. Kirstein, F., Rijpkema, S., Molkenboer, M. and Gray, J.S. (1997) The distribution and prevalence of B. burgdorferi genomospecies in Ixodes ricinus ticks in Ireland. European Journal of Epidemiology 13, 67–72. Kurtenbach, K., Dizij, A., Kampen, H., Maier, W.A., Seitz, H.M., Schaal, K.P., Moter, S.E., Kramer, M.D., Wallich,

170

L. Gern and P.-F. Humair

R., Schaible, U.E. and Simon, M.M. (1994a) The ecology of Borrelia burgdorferi in an endemic focus of Lyme borreliosis near Bonn. Applied Parasitology 35, 149–150. Kurtenbach, K., Dizij, A., Seitz, H.M., Margos, G., Moter, S.E., Kramer, M.D., Wallich, R., Schaible, U.E. and Simon, M.M. (1994b) Differential immune responses to Borrelia burgdorferi in European wild rodent species influence spirochete transmission to Ixodes ricinus L. (Acari: Ixodidae). Infection and Immunity 62, 5344–5352. Kurtenbach, K., Kampen, H., Dizij, A., Arndt, S., Seitz, H.M., Schaible, U.E. and Simon, M.M. (1995) Infestation of rodents with larval Ixodes ricinus (Acari: Ixodidae) is an important factor in the transmission cycle of Borrelia burgdorferi s.l. in German woodlands. Journal of Medical Entomology 32, 807–817. Kurtenbach, K., Carey, D., Hoodless, A.N., Nuttall, P.A. and Randolph, S.E. (1998a) Competence of pheasants as reservoirs for Lyme disease spirochetes. Journal of Medical Entomology 35, 77–81. Kurtenbach, K., Peacey, M., Rijpkema, S.G.T., Hoodless, A.N., Nuttall, P.A. and Randolph, S.E. (1998b) Differential transmission of the genospecies of Borrelia burgdorferi sensu lato by game birds and small rodents in England. Applied and Environmental Microbiology 64, 1169–1174. Kurtenbach, K., Sewell, H.S., Ogden, N.H., Randolph, S.E. and Nuttall, P.A. (1998c) Serum complement sensitivity as a key factor in Lyme disease ecology. Infection and Immunity 66, 1248–1251. Lebet, N. and Gern, L. (1994) Histological examination of Borrelia burgdorferi infections in unfed Ixodes ricinus nymphs. Experimental and Applied Acarology 18, 177–183. Le Fleche, A., Postic, D., Girardet, K., Péter, O. and Baranton, G. (1997) Characterization of Borrelia lusitaniae sp. nov. by 16S ribosomal DNA sequence

analysis. International Journal of Systematic Bacteriology 47, 921–925. Leuba-Garcia, S., Kramer, M.D., Wallich, R. and Gern, L. (1994) Characterization of Borrelia burgdorferi isolated from different organs of Ixodes ricinus ticks collected in nature. Zentralblatt für Bakteriologie 280, 468–475. Liebisch, A. and Liebisch, G. (1996) Hard ticks (Ixodidae) biting humans in Germany and their infection with Borrelia burgdorferi. In: Mitchell, R., Horn, D.J., Needham, G.N. and Welbourn, W.C. (eds) Proceedings of Acarology IX. Ohio Biological Survey, Columbus, pp. 465–467. Liebisch, A., Olbrich, S., Brand, A., Liebisch, G. and Mourettou-Kunitz, M. (1989) Natürliche Infektionen der Zeckenart Ixodes hexagonus mit Borrelien (Borrelia burgdorferi). Tierärztlische Umschau 44, 809–810. Liebisch, G., Finkbeiner-Weber, B. and Liebisch, A. (1996) The infection with Borrelia burgdorferi s.l. in the European hedgehog (Erinaceus europaeus) and its ticks. Parasitologia 38, 385. Liebisch, G., Sihns, B. and Bautsch, W. (1998a) Detection and typing of Borrelia burgdorferi sensu lato in Ixodes ricinus ticks attached to human skin by PCR. Journal of Clinical Microbiology 36, 3355–3358. Liebisch, G., Dimpfl, B., FinkbeinerWeber, B., Liebisch, A. and Frosch, M. (1998b) The red fox (Vulpes vulpes) a reservoir competent host for Borrelia burgdorferi sensu lato. In: Coons, L. and Rothschild, M. (eds) 2nd International Conference on Tick-Borne Pathogens at the Host-Vector Interface: a Global Perspective, S.I. The Conference, Kruger National Park, South Africa, p. 238. Macaigne, F. and Perez-Eid, C. (1991) Présence de Borrelia affines de B. burgdorferi chez Haemaphysalis (Alloceraea) inermis Birula, 1895 (Acarina, Ixodoidea), dans le Sud-ouest de la France. Annales de Parasitologie Humaine et Comparée 66, 269–271.

Ecology of B. burgdorferi in Europe

Marconi, R.T., Liveris, D. and Schwartz, I. (1995) Identification of novel insertion elements, restriction fragment length polymorphism patterns, and discontinuous 23S rRNA in Lyme disease spirochetes: phylogenetic analyses of rRNA genes and their intergenic spacers in Borrelia japonica sp.nov. and genomic group 21038 (Borrelia andersonii sp. nov.) isolates. Journal of Clinical Microbiology 33, 2427–2434. Marquez, F.J. and Constan, M.C. (1990) Infection d’Ixodes ricinus (L., 1758) et Haemaphysalis punctata Canestrini et Fanzago, 1877 (Acarina, Ixodidae) par Borrelia burgdorferi dans le nord de la Péninsule Ibérique (Pays Basque espagnol et Navarre). Bulletin de la Société Française de Parasitologie 8, 323–330. Masuzawa, T., Takada, N., Kudeken, M., Fukui, T., Yano, Y., Ishiguro, F., Kawamura, Y., Imai, Y. and Ezaki, T. (2001) Borrelia sinica sp. nov., a Lyme disease-related Borrelia species isolated in China. International Journal of Systematic and Evolutionary Microbiology 51, 1817–1824. Mather, T.N., Wilson, M.L., Moore, S.I., Ribeiro, J.M.C. and Spielman, A. (1989) Comparing the relative potential of rodents as reservoirs of the Lyme disease spirochete (Borrelia burgdorferi). American Journal of Epidemiology 130, 143–150. Matuschka, F.R. and Spielman, A. (1992) Loss of Lyme disease spirochetes from Ixodes ricinus ticks feeding on European blackbirds. Experimental Parasitology 74, 151–158. Matuschka, F.R., Fischer, P., Heiler, M., Richter, D. and Spielman, A. (1992) Capacity of European animals as reservoir hosts for the Lyme disease spirochete. Journal of Infectious Diseases 165, 479–483. Matuschka, F.R., Eiffert, H., Ohlenbusch, A. and Spielman, A. (1994a) Amplifying role of edible dormice in Lyme disease transmission in Central Europe. Journal of Infectious Diseases 170, 122–127.

171

Matuschka, F.R., Eiffert, H., Ohlenbusch, A., Richter, D., Schein, E. and Spielman, A. (1994b) Transmission of the agent of Lyme disease on a subtropical island. Tropical Medicine and Parasitology 45, 39–44. Matuschka, F.R., Endepols, S., Richter, D., Ohlenbusch, A., Eiffert, H. and Spielman, A. (1996) Risk of urban Lyme disease enhanced by the presence of rats. Journal of Infectious Diseases 174, 1108–1111. Matuschka, F.R., Endepols, S., Richter, D. and Spielman, A. (1997) Competence of urban rats as reservoir hosts for Lyme disease spirochetes. Journal of Medical Entomology 34, 489–493. Matuschka, F.R., Allgöwer, R., Spielman, A. and Richter, D. (1999) Characteristics of garden dormice that contribute to their capacity as reservoirs for Lyme disease spirochetes. Applied and Environmental Microbiology 65, 707–711. Matuschka, F.R., Schinkel, T.W., Klug, B., Spielman, A. and Richter, D. (2000) Relative incompetence of European rabbits for Lyme disease spirochaetes. Parasitology 121, 297–302. Mehl, R. and Traavik, T. (1983) The tick Ixodes uriae (Acari, Ixodidae) in seabird colonies in Norway. Fauna Norvegica Serie B 30, 94–107. de Michelis, S., Sewell, H.S., CollaresPereira, M., Santos-Reis, M., Schouls, L.M., Benes, V., Holmes, E.C. and Kurtenbach, K. (2000) Genetic diversity of Borrelia burgdorferi sensu lato in ticks from mainland Portugal. Journal of Clinical Microbiology 38, 2128–2133. Misonne, M.C., van Impe, G. and Hoet, P.P. (1998) Genetic heterogeneity of Borrelia burgdorferi sensu lato in Ixodes ricinus ticks collected in Belgium. Journal of Clinical Microbiology 36, 3352–3354. Miyamoto, K., Sato, Y., Okada, K., Fukunaga, M. and Sato, F. (1997) Competence of a migratory bird, redbellied thrush (Turdus chrysolaus), as an avian reservoir for the Lyme disease

172

L. Gern and P.-F. Humair

spirochetes in Japan. Acta Tropica 65, 43–51. Monin, R., Gern, L. and Aeschlimann, A. (1989) A study of the different modes of transmission of Borrelia burgdorferi by Ixodes ricinus. Zentralblatt für Bakteriologie, Suppl. 18, 14–20. Nakao, M., Miyamoto, K. and Fukunaga, M. (1994) Lyme disease spirochetes in Japan: enzootic transmission cycles in birds, rodents, and Ixodes persulcatus ticks. Journal of Infectious Diseases 170, 878–882. Nuncio, M.S., Péter, O., Alves, M.J., Bacellar, F. and Filipe, A.R. (1993) Isolamento e caracterizaçao de borrélias de Ixodes ricinus L. em Portugal. Revista Portuguesa Doenças Infecciosas 16, 175–179. Nuttall, P., Randolph, S., Carey, D., Craine, N., Livesley, A. and Gern, L. (1994) The ecology of Lyme borreliosis in the UK. In: Axford, J.S. and Rees, D.H.E. (eds) Lyme Borreliosis. Plenum Press, New York, pp. 125–129. Ogden, N.H., Nuttall, P.A. and Randolph, S.E. (1997) Natural Lyme disease cycles maintained via sheep by cofeeding ticks. Parasitology 115, 591–599. Olsen, B. (1995) Birds and Borrelia. PhD thesis, Umeå University, Umeå, Sweden. Olsen, B., Jaenson, T.G.T., Noppa, L., Bunikis, J. and Bergström, S. (1993) A Lyme borreliosis cycle in seabirds and Ixodes uriae ticks. Nature 362, 340–342. Olsen, B., Duffy, D.C., Jaenson, T.G.T., Gylfe, Å., Bonnedahl, J. and Bergström, S. (1995a) Transhemispheric exchange of Lyme disease spirochetes by seabirds. Journal of Clinical Microbiology 33, 3270–3274. Olsen, B., Jaenson, T.G.T. and Bergström, S. (1995b) Prevalence of Borrelia burgdorferi sensu lato-infected ticks on migrating birds. Applied and Environmental Microbiology 61, 3082–3087. Perret, J.L., Guigoz, E., Rais, O. and Gern, L. (2000) Influence of saturation deficit and temperature on Ixodes ricinus tick questing activity in a Lyme borreliosis-

endemic area (Switzerland). Parasitology Research 86, 554–557. Péter, O. and Bretz, A.G. (1992) Polymorphism of outer surface proteins of Borrelia burgdorferi as a tool for the classification. Zentralblatt für Bakteriologie 277, 28–33. Péter, O., Bretz, A.G. and Bee, D. (1995) Occurrence of different genospecies of Borrelia burgdorferi sensu lato in ixodid ticks of Valais, Switzerland. European Journal of Epidemiology 11, 463–467. Petney, T.N., Hassler, D., Bröckner, M. and Maiwald, M. (1996) Comparison of urinary bladder and ear biopsy samples for determining prevalence of Borrelia burgdorferi in rodents in central Europe. Journal of Clinical Microbiology 34, 1310–1312. Pichon, B., Godfroid, E., Hoyois, B., Bollen, A., Rodhain, F. and Perez, C. (1995) Simultaneous infection of Ixodes ricinus by two Borrelia burgdorferi sensu lato species: possible implications for clinical manifestations. Emerging Infectious Diseases 1, 89–90. Picken, R.N., Cheng, Y., Strle, F. and Picken, M.M. (1996) Patient isolates of Borrelia burgdorferi sensu lato with genotypic and phenotypic similarities to strain 25015. Journal of Infectious Diseases 174, 1112–1115. Postic, D., Assous, M.V., Grimont, P.A.D. and Baranton, G. (1994) Diversity of Borrelia burgdorferi sensu lato evidenced by restriction fragment length polymorphism of rrf (5S)–rrl (23S) intergenic spacer amplicons. International Journal of Systematic Bacteriology 44, 743–752. Postic, D., Korenberg, E., Gorelova, N., Kovalevski, Y.V., Bellenger, E. and Baranton G. (1997) Borrelia burgdorferi sensu lato in Russia and neighbouring countries: high incidence of mixed isolates. Research in Microbiology 148, 691–702. Postic, D., Marti Ras, N., Lane, R.S., Hendson, M. and Baranton, G. (1998) Expanded diversity among Californian Borrelia isolates and description of

Ecology of B. burgdorferi in Europe

Borrelia bissettii sp.nov. (formerly Borrelia group 127). Journal of Clinical Microbiology 36, 3497–3504. Randolph, S.E. and Craine, N.G. (1995) General framework for comparative quantitative studies on transmission of tick-borne diseases using Lyme borreliosis in Europe as an example. Journal of Medical Entomology 32, 765–777. Randolph, S. and Storey, K. (1999) Impact of microclimate on immature tick–rodent host interactions (Acari: Ixodidae): implications for parasite transmission. Journal of Medical Entomology 36, 741–748. Randolph, S.E., Gern, L. and Nuttall, P.A. (1996) Co-feeding ticks: epidemiological significance for tick-borne pathogen transmission. Parasitology Today 12, 472–479. Richter, D., Endepols, S., Ohlenbusch, A., Eiffert, H., Spielman, A and Matuschka, F.R. (1999) Genospecies diversity of Lyme disease spirochetes in rodent reservoirs. Emerging Infectious Diseases 5, 291–296. Richter, D., Spielman, A., Komar, N. and Matuschka, F.R. (2000) Competence of American robins as reservoir hosts for Lyme disease spirochetes. Emerging Infectious Diseases 6, 133–138. Rijpkema, S.G.T., Molkenboer, M.J.C.H., Schouls, L.M., Jongejan, F. and Schellekens, J.F.P. (1995) Simultaneous detection and genotyping of three genomic groups of Borrelia burgdorferi sensu lato in Dutch Ixodes ricinus ticks by characterization of the amplified intergenic spacer region between 5S and 23S rRNA genes. Journal of Clinical Microbiology 33, 3091–3095. Rijpkema, S., Golubic, D., Molkenboer, M., Verbeek-De Kruif, N. and Schellekens, J. (1996) Identification of four genomic groups of Borrelia burgdorferi sensu lato in Ixodes ricinus ticks collected in a Lyme borreliosis endemic region of northern Croatia. Experimental and Applied Acarology 20, 23–30.

173

Saint Girons, I., Gern, L., Gray, J.S., Guy, E.C., Korenberg, E., Nuttall, P.A., Rijpkema, S.G.T., Schönberg, A., Stanek, G. and Postic, D. (1998) Identification of Borrelia burgdorferi sensu lato species in Europe. Zentralblatt für Bakteriologie 287, 190–195. Stanek, G. and Simeoni, J. (1989) Are pigeons’ ticks transmitters of Borrelia burgdorferi to humans? Zentralblatt für Bakteriologie Suppl. 18, 42–43. Strle, F., Picken, R.N., Cheng, Y., Cimperman, J., Maraspin, V., LotricFurlan, S., Ruzic-Sabljic, E. and Picken, M.M. (1997) Clinical findings for patients with Lyme borreliosis caused by Borrelia burgdorferi sensu lato with genotypic and phenotypic similarities to strain 25015. Clinical Infectious Diseases 25, 273–280. Tälleklint, L. (1996) Lyme borreliosis spirochetes in Ixodes ricinus and Haemaphysalis punctata ticks (Acari: Ixodidae) on three islands in the Baltic Sea. Experimental and Applied Acarology 20, 467–476. Tälleklint, L. and Jaenson, T.G.T. (1993) Maintenance by hares of European Borrelia burgdorferi in ecosystems without rodents. Journal of Medical Entomology 30, 273–276. Tälleklint, L. and Jaenson, T.G.T. (1994) Transmission of Borrelia burgdorferi s.l. from mammal reservoirs to the primary vector of Lyme borreliosis, Ixodes ricinus (Acari: Ixodidae), in Sweden. Journal of Medical Entomology 31, 880–886. Toutoungi, L. and Gern, L. (1993) Ability of transovarially and subsequent transstadially infected Ixodes hexagonus ticks to maintain and transmit Borrelia burgdorferi in the laboratory. Experimental and Applied Acarology 17, 581–586. Toutoungi, L.N., Gern, L., Aeschlimann, A. and Debrot, S. (1991) A propos du genre Pholeoixodes, parasite des carnivores en Suisse. Acarologia XXXII, 311–328. Wang, G., van Dam, A.P., Le Fleche, A., Postic, D., Péter, O., Baranton, G., de

174

L. Gern and P.-F. Humair

Boer, R., Spanjaard, L. and Dankert, J. (1997) Genetic and phenotypic analysis of Borrelia valaisiana sp. nov. (Borrelia genomic groups VS116 and M19). International Journal of Systematic Bacteriology 47, 926–932. Wang, G.Q., van Dam, A.P. and Dankert, J. (1999) Phenotypic and genetic characterization of a novel Borrelia burgdorferi sensu lato isolate from a patient with Lyme borreliosis. Journal of Clinical Microbiology 37, 3025–3028. Will, G., Jauris-Heipke, S., Schwab, E., Busch, U., Rössler, D., Soutschek, E., Wilske, B. and Preac-Mursic, V. (1995) Sequence analysis of ospA genes shows homogeneity within Borrelia burgdorferi sensu stricto and Borrelia afzelii strains but reveals major subgroups within the Borrelia garinii species. Medical Microbiology and Immunology 184, 73–80. Wilske, B., Preac-Mursic, V., Göbel, U.B., Graf, B., Jauris, S., Soutschek, E., Schwab, E. and Zumstein, G. (1993) An OspA serotyping system for Borrelia

burgdorferi based on reactivity with monoclonal antibodies and OspA sequence analysis. Journal of Clinical Microbiology 31, 340–350. Wilske, B., Busch, U., Eiffert, H., Fingerle, V., Pfister, H.W., Rössler, D. and Preac-Mursic, V. (1996) Diversity of OspA and OspC among cerebrospinal fluid isolates of Borrelia burgdorferi sensu lato from patients with neuroborreliosis in Germany. Medical Microbiology and Immunology 184, 195–201. Younsi, H., Postic, D., Baranton, G. and Bouattour, A. (2001) High prevalence of Borrelia lusitaniae in Ixodes ricinus ticks in Tunisia. European Journal of Epidemiology 17, 53–56. Zhioua, E., Bouattour, A., Hu, C.M., Gharbi, M., Aeschlimann, A., Ginsberg, H. and Gern, L. (1999) Infections of Ixodes ricinus (Acari: Ixodidae) by Borrelia burgdorferi sensu lato in North Africa (Tunisia). Journal of Medical Entomology 36, 216–218.

Ecology of Borrelia burgdorferi sensu lato in Russia

7

Edward I. Korenberg, Natalya B. Gorelova, and Yurii V. Kovalevskii Gamaleya Research Institute for Epidemiology and Microbiology, Russian Academy of Medical Sciences, 18 Gamaleya Street, Moscow, 123098 Russia

Introduction Purposeful investigations of borrelioses transmitted by hard ticks began in Russia in 1984, immediately after the corresponding spirochaetes were found in European Ixodes ricinus ticks (Burgdorfer et al., 1983) and the aetiological agent of Lyme disease was defined as a new species, Borrelia burgdorferi sp. nov. (Johnson et al., 1984). However, Russian dermatologists had described clinical manifestations of the disease under different names – atrophia cutis maculata, atrophia cutis idiopathica progressiva acquisita, idiopathic acquired skin atrophy, or autonomous skin atrophy – as early as the turn of the 19th and 20th centuries (Nikol’skii, 1896; Mescherskii, 1898; Pisemskii, 1902; Pospelov, 1905, 1906). Nearly 25 years before the causes of the disease were revealed, Lisovskaya (1958) characterized in detail the clinical picture of chronic atrophic acarodermatitis in her dissertation. Moreover, since the 1940s, neuropathologists have devoted special attention to the diseases that occurred after tick bites in the regions where tick-borne encephalitis (TBE) was endemic and which were accompanied by extensive ‘acaroerythema’ or ‘annular acaroerythema’ (as it was called at that time) and various neurological disturbances, and where serological tests for TBE were negative. Such cases were recorded in the northwestern, central and eastern areas of European Russia, in the Urals, West Siberia and the Russian Far East. In the 1960s, several specialists in clinical medicine concluded that these were aetiologically independent diseases unrelated to TBE. The detailed descriptions of their general clinical picture and of migratory erythema observed after tick bites provide evidence so that many of these cases can be diagnosed retrospectively as Lyme disease (LD) or Lyme borreliosis (LB). Both terms actually refer to the whole group of aetiologically independent diseases. Their pathogens © 2002 CAB International. Lyme Borreliosis: Biology, Epidemiology and Control (eds J. Gray, O. Kahl, R.S. Lane and G. Stanek)

175

176

E.I. Korenberg et al.

have become associated ecologically with certain ticks of the genus Ixodes. It was proposed to name this group ixodid tick-borne borrelioses (ITBB) in contrast to argasid tick-borne borrelioses (ATBB) associated with soft ticks (Korenberg, 1993, 1995, 1996; Korenberg and Kryuchechnikov, 1996). In Russia, the cases of ITBB were first identified serologically in 1985 (Korenberg et al., 1986). Patients with these infections have been recorded in 53 administrative territories of the Russian Federation, from the Baltic Sea to the Far East and southern Sakhalin. Thus, much of the worldwide range of B. burgdorferi s.l. is on Russian territory (Korenberg et al., 1987b; Korenberg, 1994, 1995). As diagnostics improve, ITBB morbidity is steadily approaching the predicted level of 10,000–12,000 fresh cases per year, and in some years this level may be even higher (Korenberg, 1995). Thus, the total numbers of ITBB cases in Russia in 1998 and 1999 were 8606 and 8470, or 5.9 and 5.8 cases per 100,000 people, respectively. These infections are now amongst the foremost causes of morbidity within the group of zoonoses with natural focality, with the greatest number of cases being recorded in the Urals, West Siberia and the Volga–Vyatka region. Studies on natural ITBB foci in Russia began in the northwest and the far east (Korenberg et al., 1987a, 1988, 1989, 1991a,b). Those results provided the basis for the following principal conclusions: (i) the tick Ixodes persulcatus Schultze is the main vector of borreliae in the greater part of the ITBB range; (ii) the most active foci are found in broad-leaved, mixed and southern taiga forests; (iii) the tick I. ricinus (L.) is a borrelia vector in the European part of Russia, but is less efficient than I. persulcatus; and (iv) the distribution of ITBB and the basic features of their epidemiology in Russia are similar to those of TBE. These conclusions were fully confirmed in subsequent studies (Korenberg, 1994; Naumov, 1999). These two tick species were the source of the first B. burgdorferi s.l. isolates obtained and serotyped in Russia (Kryuchechnikov et al., 1988, 1990, 1993). Some of them (Ip 3, Ip 21, Ip 89, Ip 90, Ir 210 and others) were used by specialists from several countries in studies on the genetic structure, taxonomy and phylogeny of Borrelia spp. (Korenberg and Kryuchechnikov, 1996), including characterization of their genogroups and original description of individual genospecies (for example, Baranton et al., 1992; Belfaiza et al., 1993; Fukunaga et al., 1993; Assous et al., 1994; Postic et al., 1994; Liveris et al., 1995). The data accumulated in studies on the main aspects of B. burgdorferi s.l. ecology in Russia are presented below.

Vectors and B. burgdorferi Genospecies Present Isolates of borreliae were obtained from three tick species: I. persulcatus, I. ricinus, and Ixodes (Exopalpiger) trianguliceps. However, attempts to isolate borreliae from Dermacentor reticulatus ticks collected in northwestern Russia and from Haemaphysalis concinna and Haemaphysalis japonica ticks from the far east of Russia were not successful (Korenberg et al., 1997). Information concerning borrelial infection in

Ecology of B. burgdorferi in Russia

177

D. reticulatus, Dermacentor marginatus, Dermacentor nuttalii, H. concinna and H. japonica ticks in West Siberia and the Far East is based on the results of dark-field microscopy only (Ivanov et al., 1996; Rudakova and Matuschenko, 1996) and has not been confirmed by more objective microbiological methods. The taiga tick I. persulcatus and the forest tick I. ricinus are the main vectors of borreliae in natural foci and the source of these pathogens for man. The taiga tick is the vector in the southern forest zone of Eurasia, the vast area extending from the Far East to the Russian northwest (beyond this zone, its distribution is limited mainly to some Asian countries), and the forest tick is the vector in similar forests of Europe. In much of European Russia, where these ticks are sympatric (Figs 7.1 and 7.2), natural foci with the two vector species exist, as in the case of TBE (Korenberg, 1994; Filippova, 1999). Indices of borrelial prevalence in adult I. persulcatus and I. ricinus ticks in different Russian regions, as determined by dark-field microscopy, are shown in Table 7.1. Prevalence is high everywhere, but it appears that in general slightly higher indices are characteristic of the Ural, West Siberian and, probably, the Far-East regions. The prevalence of borreliae in I. ricinus ticks is markedly lower than in I. persulcatus, both in individual regions, for example Leningrad and Moscow oblasts (a large Russian administrative area), and in general. In this connection it should be noted that the presence and proportion of one vector in an ecosystem does not have any significant effect on the

Fig. 7.1. Distribution of two Borrelia garinii variants in different regions of Russia (according to Postic et al., 1997, with changes and supplements). (1) St Petersburg; (2) Yaroslavl; (3) Perm; (4) Sverdlovsk; (5) Kurgan; (6) Tyumen; (7) Novosibirsk; (8) Gornii Altai Republic; (9) Krasnoyarsk; (10) Khabarovsk; (11) Primor’e; (12) Sakhalin; (13) Kaliningrad; (14) Krasnodar; (15) Stavropol; (16) Kazakhstan.

178

E.I. Korenberg et al.

Fig. 7.2. Distribution of Borrelia afzelii, Borrelia burgdorferi sensu stricto, Borrelia lusitaniae and Borrelia valaisiana in Russia (according to Postic et al., 1997, with changes and supplements). (1) St Petersburg; (2) Yaroslavl; (3) Perm; (4) Sverdlovsk; (5) Kurgan; (6) Tyumen; (7) Novosibirsk; (8) Gornii Altai Republic; (9) Krasnoyarsk; (10) Khabarovsk; (11) Primor’e; (12) Sakhalin; (13) Kaliningrad; (14) Krasnodar; (15) Stavropol.

extent of infection and on the epizootic and epidemiological significance of the other vector (Korenberg et al., 2001a). The tick I. trianguliceps, widespread in the forested zone of Eurasia from Great Britain to the southern end of the Baikal Lake in Russia, does not attack people and feeds mainly on small mammals (Korenberg and Lebedeva, 1969; Filippova, 1977). According to data on borrelial cultivation, the proportion of infected larvae and nymphs is about 10% (Gorelova et al., 1996); according to the results of immunofluorescence assay (IFA) with monoclonal antibodies, the prevalence of infection is far higher (Grigoryeva and Tretyakov, 1998). This vector may be of some epizootiological significance (Gorelova et al., 1996; Korenberg et al., 1997), but it remains unclear whether I. trianguliceps can maintain borrelial circulation in the absence of I. persulcatus and I. ricinus ticks. The tick Ixodes pavlovskyi, whose contemporary disjunctive range consists of the western (Altai–Sayan) and Far Eastern parts, is closely related to I. persulcatus, both morphologically and ecologically, and is sympatric with the latter virtually everywhere (Filippova, 1977, 1999). Borreliae were first isolated from I. pavlovskyi ticks collected in the southwestern Altai (Fig. 7.1, No. 16), near the Kazakh–Russian border (Gorelova et al., 2001a). The role of this species in ITBB epizootiology and epidemiology requires further investigation. Information about B. burgdorferi genospecies present in different species of vec-

Table 7.1. Prevalence of Borrelia burgdorferi sensu lato in unfed adult ixodid ticks in different regions of Russia (microscopy data). Prevalence of Borrelia in ticks

Years

Number of ticks studied

Average

Min.–Max.

Reference

1995–1997 1986–1992 1995–1997 1990

1868 4059 1559 210

12.0 33.2 27.0 29.5

8.4–19.8 8.6–47.8 24.2–30.7 20.5–39.1

1990–1991

484

29.8

20.0–36.5

Rybakova and Evsyukova, 1999 Korenberg et al., 2001a Dubinina and Alekseev, 1999 Korenberg et al., 2001a (with supplements) Korenberg et al., 2001a (with supplements)

Central part of European Russia Moscow oblast

1990–1991

174

28.7

20.0–48.0

Alekseev, 1993

Ural region Perm oblast Perm city

1993–2000 1993–1998

1324 1311

42.7 27.2

29.5–60.8 18.4–35.6

Sverdlovsk oblast Chelyabinsk oblast Kurgan oblast

1993–1995 1994–1997 1991–1995

228 502 759

37 8.0 22.8

31–42 1.1–18.6 4–45

Korenberg et al., 1999 (with supplements) Korenberg et al., 2001b (with supplements) Belyaeva, 1998 Tarasov et al., 2000 Smirnova et al., 1996

West Siberia region Tyumen oblast Omsk oblast Novosibirsk oblast Altai territory

1990–1994 1991–1995 1991–1995 1991–1995

2046 900 292 248

35.6 25.0 26.4 21.8

22.6–44.6 11.9–31.3 13.3–58.3 17.0–46.1

Regions/provinces Ixodes persulcatus

Novgorod oblast

179

Kolchanova, 1997 Rudakova and Matuschenko, 1996 Rudakova and Matuschenko, 1996 Rudakova and Matuschenko, 1996

Ecology of B. burgdorferi in Russia

Northern and Northwest regions Vologda oblast Leningrad oblast St Petersburg city Pskov oblast

Table 7.1. Continued 180

Prevalence of Borrelia in ticks Average

Min.–Max.

Reference

1993–1995

3036

22.1

12.2–30.6

1992–1995 ~ 2500 1994–1995 448

19 26.8

12–58 25.0–28.6

Naumov et al., 1994 Vasilieva and Naumov, 1996 Gorin et al., 1996 Apostolova et al., 1996

1986–1987 1995 1995

288 99 183

22.5 41.4a 25.1a

21.2–22.9 – –

Korenberg et al., 1989 Sato et al., 1996 Sato et al., 1996

Northwestern region Kaliningrad oblast Leningrad oblast Novgorod oblast

1995–1998 1986–1992 1990–1991

2470 2559 38

11.1 16.7 21.0

10.0–13.6 5.0–21.9 6.3–37.5

Dubinina and Alekseev, 1999 Korenberg et al., 2001a Korenberg et al., 2001a (with supplements)

Central part of European Russia Kaluga oblast Moscow oblast

1992–1998 1991–1992

1015 308

15.5 18.5

10.9–19.8 5.6–35.2

Burenkova, 2000 Alekseev, 1993

North Caucasus region Stavropol territory Krasnodar territory

1996–1997 2000

207 112

15.9a 18.8a

– –

Regions/provinces East Siberia region Khakassia Republic Irkutsk oblast Irkutsk city Far Eastern region Khabarovsk territory Primorye territory Sakhalin oblast Ixodes ricinus

a

Ticks were studied by culturing of their internal organs on BSK medium. –, no data.

Korenberg et al., 1999 Gorelova et al., 2001b

E.I. Korenberg et al.

Years

Number of ticks studied

Ecology of B. burgdorferi in Russia

181

tors and reservoir hosts in different Russian regions is based on the results of identification of isolates obtained in several studies (Sato et al., 1996; Korenberg et al., 1997; Masuzawa et al., 1997; Postic et al., 1997; Saint Girons et al., 1998), and on unpublished data obtained in the past few years. More than 800 isolates have been identified so far and are now kept in one of the largest collections of B. burgdorferi s.l. isolates in the world in the Museum of Borrelia, Laboratory of Vectors of Infections (Gamaleya Research Institute for Epidemiology and Microbiology, Moscow) (Gorelova, 1993). In Russia, the spirochaetes B. garinii and B. afzelii occur virtually everywhere and jointly circulate in natural foci (Figs 7.1 and 7.2). They were isolated from three tick species (Table 7.2). The genospecies B. garinii is represented by two types, 20047T and NT29. The former type occurs in all the vectors, whereas the latter has been found only in I. persulcatus and I. trianguliceps. It may be that B. garinii NT29 has not been found in I. pavlovskyi simply because only four borrelial isolates have as yet been obtained from this tick species (Gorelova et al., 2001a). However, in the case of I. ricinus, 79 spirochaetal isolates from different Russian regions have been typed to date, and the absence of B. garinii NT29 among them is difficult to explain in the same way. Thus, we conclude that the distribution of B. garinii NT29 is associated with that of I. persulcatus ticks (Postic et al., 1997). The species B. valaisiana, B. lusitaniae and B. burgdorferi s.s. have been isolated only from I. ricinus ticks in the European part of Russia, with the proportion of ticks infected by B. burgdorferi s.s. being less than 1% (Korenberg et al., 1999; Table 7.2. Genospecies of Borrelia burgdorferi sensu lato isolated in Russia from vectors and reservoir hosts. Species of vectors and reservoir hosts

B. garinii B. burgdorferi sensu B. Variant Variant B. afzelii 20047T NT29 stricto lusitaniae

Ixodid ticks Ixodes persulcatus Ixodes ricinus Ixodes trianguliceps Ixodes pavlovskyi Small mammals Clethrionomys glareolus Clethrionomys rufocanus Clethrionomys rutilus Microtus oeconomus Microtus agrestis Apodemus uralensis Apodemus peninsulae Sorex araneus

+ + +

+ + + +

+

+ + + + + +

+ + + +

+ + + +

Blank spaces indicate species not found.

+ +

+ + +

+

B. valaisiana

+

182

E.I. Korenberg et al.

Gorelova et al., 2001b) although it cannot be excluded that in some localities the prevalence of this recognized human pathogen in I. ricinus is higher. According to Semenov et al. (2000), DNA of this spirochaete genospecies was detected by PCR in I. ricinus and I. persulcatus ticks from the northwest region of Russia. These findings need to be verified by isolation of the pathogen. In any case, B. burgdorferi s.s. is of limited significance as an aetiological agent of ITBB in Russia, and the aetiology, epidemiology and clinical manifestations of these diseases are mainly attributable to B. garinii and B. afzelii (Korenberg, 1995). Different combinations of mixed isolates have been obtained from three tick species collected in many Russian regions (Table 7.3). For example, according to our long-term data on borrelial culturing, the proportions of mixed-infected I. persulcatus in the southern taiga forests of the Ural region are 1.6 ± 0.9% among adult ticks collected from the vegetation and 1.5 ± 1.1% among nymphs. Furthermore, mixed infection of I. persulcatus and I. ricinus ticks from the northwest was revealed by PCR analysis (Dubinina and Alekseev, 1999). These data support the conclusion that practically all possible combinations of mixed infections with two or more Borrelia spp. circulating within a certain area can occur Table 7.3. Distribution of different Borrelia burgdorferi s.l. mixed isolates from vectors and reservoir hosts in Russia. Variants of mixed isolates

Species of vectors and reservoir hosts Ixodid ticks Ixodes persulcatus

Region

B. garinii NT29 + B. afzelii

Central part of European Russia Ural region West Siberian region Far Eastern region North Caucasian region Ural region

Ixodes ricinus Ixodes trianguliceps Small mammals Clethrionomys glareolus Ural region Clethrionomys rutilus Ural region Microtus oeconomus Ural region Clethrionomys rufocanus and Apodemus peninsulae Far Eastern region

+ + +

B. garinii NT29 + B. garinii 20047T

B. garinii NT29 + B. garinii B. garinii B. garinii 20047T + 20047T + 20047T+ B. afzelii B. valaisiana B. afzelii

+

+ + +

+

+

+ +

+

+

+

+

+

+

+ + +

+

+

+?

+?

+?

+?, According to Masuzawa et al., 1997. Particular variant not indicated. Blank spaces indicate species not recorded.

Ecology of B. burgdorferi in Russia

183

in both unfed and engorged (or partially engorged) ticks at any phase of their development (Postic et al., 1997; Korenberg, 1999).

Host Preferences of Ticks The list of vertebrates that can serve as hosts of I. ricinus and I. persulcatus comprises at least 300 species and includes virtually all terrestrial mammals occurring within the ranges of these vectors, many ground-foraging bird species and even reptiles and amphibians, which, however, have no apparent significance in this respect (Filippova, 1977; Naumov, 1985). Larvae and nymphs parasitize more than 50 species of small mammals. The roles of these animals as hosts of larvae and nymphs are fairly similar, but insectivores are much less attractive for nymphs. Throughout the forest zone of Russia, voles of the genus Clethrionomys and shrews of the genus Sorex are of special significance as tick hosts. This universal Clethrionomys–Sorex type of host consists of two variants (subtypes). The first one is characteristic of European Russia (and a large part of central Europe), where most immature ticks feed on bank voles (Clethrionomys glareolus) and common shrews (Sorex araneus). The second variant is characteristic of Asian Russia, where the genus Sorex retains its role but the bank vole is replaced by the ruddy vole (Clethrionomys rutilus) and the grey large-toothed red-backed vole (Clethrionomys rufocanus) (Korenberg, 1979; Tupikova et al., 1980). In addition to small mammals, larger animals, such as hedgehogs, squirrels and, in Asian Russia, chipmunks, are readily attacked by larvae and nymphs. The average number of larvae and nymphs per animal can vary widely, ranging in different regions from 0.5 to 69.7 in the case of the red squirrel (Sciurus vulgaris) and from 0.5 to 29.3 in the case of the Siberian chipmunk (Tamius sibiricus) (Korenberg, 1979). Nymphs also frequently attack wild hoofed animals. Larvae of I. ricinus and I. persulcatus very rarely attack large animals in Russia. Immature I. persulcatus and I. ricinus ticks have been found on more than 180 species of bird. However, the number of these bird species is greater in the Far East, within the optimum I. persulcatus range; it gradually decreases in the eastern and central regions of European Russia and increases again in the regions located farther to the west, within the I. ricinus optimum range (Korenberg, 1962). The only birds attacked by larvae and nymphs throughout this area are the forest species of grouse (Tetraonidae spp.), thrushes (Turdus spp.), pipits (Anthus spp.) and buntings (Emberiza spp.). It appears that, as in the case of TBE (Korenberg, 1964), the involvement of birds in the circulation of Borrelia spp. in most Russian regions is a relatively rare phenomenon (see below). Hares are of special significance as hosts because ticks parasitize them at all three phases of tick development: for example, on average one mountain hare (Lepus timidus) can be simultaneously infested by up to 44 larvae, 24 nymphs and 85 adult I. persulcatus (Mishin, 1964). Adult ticks also attack virtually all larger animals, both wild and domestic, and the significance of different groups of hosts in a certain region depends on the numbers of livestock, conditions of their

184

E.I. Korenberg et al.

maintenance and pasturing, and the abundance of large wild mammals. In the southern taiga forests of the eastern East European Plain, for example, cattle are parasitized by approximately 52% of adult I. persulcatus ticks, hares (L. timidus) 34%, elks (Alces alces) 4%, red foxes (Vulpes vulpes) 2% and other species of large mammals 8% (Korenberg, 1978). In populated areas of the Far East, approximately 62% of I. persulcatus ticks feed on cattle, 3% on horses, 1% on dogs, 24% on red deer (Cervus elaphus), 3% on roe deer (Capreolus capreolus), 6% on hares (L. timidus) and only about 1% on other animal species (Savitskii, 1970). It should be noted that in contrast to I. ricinus ticks, which attack people as both nymphs and adults, I. persulcatus nymphs very rarely attack people. Larvae of both species do not as a rule attack people in Russia (Korenberg and Kovalevskii, 1981; Korenberg, 1994; Korenberg et al., 2001b). As in the case of TBE, a very important factor affecting the development of the epizootic process in natural ITBB foci is that ticks of different developmental stages, belonging to different generations, often parasitize the same host. This situation provides an especially efficient horizontal pathway of pathogen transmission, bringing together the vector generations of at least 5 successive years (Korenberg and Kovalevskii, 1994, 1999).

Host Preferences of Spirochaetes Eleven species of small mammals from various Russian regions have been analysed for borrelial infection, but isolates were obtained and identified from only eight species (Tables 7.2 and 7.4). In all these regions animals proved to contain the same predominant Borrelia genospecies and their combinations as those found in ixodid ticks (Table 7.3), with B. afzelii and both variants of B. garinii having been isolated from all studied species of red-backed voles (Clethrionomys spp.). No preference of a certain genospecies for a particular host species was revealed, and the absence of some Borrelia spp. from certain species of small mammals (Table 7.4) may be due to an insufficient amount of data. In addition to the species listed in Table 7.4, we examined 13 species of small to medium-sized mammals (Microtus arvalis, Mus musculus, Sicista betulina, Sorex caecutiens, Sorex isodon, Sorex minutus, Neomys fodiens, Talpa europaea, Erinaceus europaeus, Arvicola terrestris, Cricetus cricetus, L. timidus and Mustela nivalis) inhabiting an active ITBB focus in the Ural region (see below), for the presence of Borrelia spp. in the organs. All the 164 cultures in BSKII medium (from urinary bladder, heart, spleen, kidney and, for some species, ear skin biopsies) were negative. Our study on the presence of borreliae in birds, which is still the only work of its kind in Russia, was performed in the same natural focus in 1998. Culture medium was inoculated with 94 skin biopsies (taken from the head near the ear opening) and brain samples of 47 birds belonging to four species closely associated with the lower forest layer, namely, 34 Turdus pilaris, nine Turdus musicus, two Emberiza citrinella and two Anthus trivialis. No borreliae were found in any of these cultures, which is evidence that birds may not be involved in the circula-

Ecology of B. burgdorferi in Russia

185

Table 7.4. Distribution of different Borrelia burgdorferi sensu lato genospecies in small mammals in Russia.

+ + + +

+ + + + + +

+ + + + + +

+ + + + + + + + +

+

Grigoryeva and Tretyakov, 1998a

Grigoryeva, 1996a Gorelova et al., 1995 (with supplements); Korenberg et al., 1997 (with supplements)

Matuschenko et al., 1996

  

+ + + + +

        

North West region Clethrionomys glareolus Clethrionomys rutilus Microtus arvalis Apodemus uralensis Sorex araneus Ural region Clethrionomys glareolus Clethrionomys rutilus Clethrionomys rufocanus Microtus oeconomus Microtus agrestis Apodemus uralensis Sorex araneus West Siberian region Clethrionomys glareolus Clethrionomys rutilus Clethrionomys rufocanus Microtus oeconomus Apodemus agrarius Micromys minutus Sorex sp. Far East region Clethrionomys rufocanus Apodemus peninsulae

B. burgdorferi sensu lato Reference

      

B.garinii 20047T NT29 B. afzelii

    

Species of small mammals

Sato et al., 1996

a

Borreliae were revealed by stained blood smears and imprints of kidney cuts, and also by silver-staining of histological preparations. Blank spaces indicate species not found.

tion of Borrelia spp., at least in the study area, though it is also possible that some bird spirochaetes are particularly difficult to culture. In regions where birds are heavily parasitized by ticks, for example the Far East of Russia (see Miyamoto and Masozuwa, Chapter 8), some bird species appear be reservoir hosts of Borrelia spp.

Reservoir Hosts More detailed data on the reservoir hosts and vectors of different Borrelia genospecies were obtained during long-term studies (1992–2000) on the natural ITBB foci located in the east European southern taiga forests of the Ural region.

186

E.I. Korenberg et al.

Altogether 595 borrelial isolates were obtained and identified from seven species of small mammals and from I. persulcatus and I. trianguliceps ticks of all developmental stages. Some researchers are of the opinion that borreliae persisting in the internal organs of small mammals are of different genospecies from those in the skin. In particular, it is maintained that B. garinii occurs exclusively or mainly in the internal organs, but is absent from the skin and, hence, is not transmitted to the ticks that feed on these animals. On this basis, the conclusion is drawn that small mammals do not serve as reservoir hosts of B. garinii and can transmit only B. afzelii to ticks (Humair et al., 1995, 1999; Gern et al., 1997; Hu et al., 1997; Kurtenbach et al., 1998). However, most of the isolates we obtained from ear skin biopsies of small mammals as well as from I. persulcatus larvae removed from these animals were identified as two variants of B. garinii; moreover, the incidence of these spirochaetes in the internal organs (the urinary bladder) was even slightly lower than in the skin (Table 7.5). Thus, small mammals proved to be good reservoir hosts not only for B. afzelii, but also for both these particular B. garinii genospecies variants. The absence of any fundamental differences in the composition of Borrelia spp. in the internal organs and skin biopsies of small mammals makes it possible to assess the aetiological structure of natural foci by analysing all the isolates obtained (Fig. 7.3). The most abundant rodent species in the study region is the bank vole C. glareolus, which accounts for more than 70% of animals captured. More than 40% of Borrelia spp. isolates obtained from bank voles belonged to B. garinii NT29, approximately 25% to B. afzelii, less than 20% to B. garinii 20047T, and the rest were mixed isolates (Fig. 7.3, histogram 2). Approximately the same ratio of B. garinii (both variants) to B. afzelii was observed among isolates obtained from the other six species of reservoir hosts (Fig. 7.3, histogram 3). Moreover, this ratio was also characteristic of partially or fully engorged immature I. persulcatus removed from the animals and of unfed nymphs and adult ticks collected from the vegetation. On the whole, this is evidence for the weakness or even the absence of directional selection for a certain Borrelia genospecies at different links of the enzootic cycle. On the other hand, some differences in the relation of Borrelia spp. to their associated hosts and ticks were evident, though further study is needed to verify them. Two groups of facts are relevant in this context. Firstly, B. garinii 20047T occurred more frequently in all three species of Clethrionomys voles, whereas none of the 35 unmixed isolates from the meadow voles of the genus Microtus was identified as B. garinii 20047T; this variant was only found in a few mixed isolates in these voles. Meadow voles may therefore be less efficient as reservoir hosts of B. garinii 20047T than the red-backed voles (Clethrionomys spp.). Secondly, B. afzelii occurred much less frequently in fully or partially engorged I. trianguliceps ticks removed from small mammals (Fig. 7.3, histogram 7) than in the animals proper or in immature and adult I. persulcatus ticks. This may be evidence that I. trianguliceps is less susceptible to B. afzelii than to B. garinii and, therefore, there is some selection of one of the Borrelia genospecies by an accessory vector.

B. garinii Total

Source

No. of isolates identified

Absolute %

Consisting of (absolute number): Mixture Variant Variant 20047T + NT29 NT29 20047T

B. afzelii

Absolute

Mixture B. garinii + B. afzelii

%

Absolute

%

Small mammals Ear skin biopsies 66 Urinary bladder 120

56 65

84.9 54.2

15 10

36 48

5 7

8 46

12.1 38.3

2 9

3.0 7.5

Ticks Larvae

30

75.0

16

11

3

8

20.0

2

5.0

40

Ecology of B. burgdorferi in Russia

Table 7.5. Results of identification of Borrelia burgdorferi sensu lato isolates, obtained in the Ural region from small mammals and I. persulcatus ticks collected from them (1992–2000).

187

20

40

60

0

Incidence of species (%)

20

4

3

40

60

0

20

40

60

Incidence of species (%)

Incidence of species (%)

6

7

5

0

20

40

60

Incidence of species (%)

A B C D

0

20

40

60

0

20

40

60

0

20

40

60

E.I. Korenberg et al.

0

2

188

1

Ecology of B. burgdorferi in Russia

189

Ixodid Tick-borne Borrelioses Biotopes As in the case of TBE (Korenberg and Kovalevskii, 1981, 1999), the range of ITBB in Russia coincides with the ranges of I. ricinus and I. persulcatus ticks (Korenberg, 1994). Within the ITBB range, large areas are free from natural foci, because of the absence of landscape-biocenotic conditions necessary for the existence of these vectors. The mosaic of ITBB biotopes is wholly determined by specific features of the biotopic preference of the main vectors, which varies in different climatic zones and landscapes because the ticks are always confined to the biotopes with hygrothermal conditions appropriate for them. The patchwork pattern of ITBB distribution is eventually formed by the effects of typical zonal, intrazonal and extrazonal conditions and their combinations. The tick I. ricinus usually prefers moderately hygrophilic and mesophilic biotopes in plain and mountain broad-leaved and mixed forests. In the southern and (in places) mid-taiga subzones, it occurs in warmer areas, such as mixed coniferous and deciduous and light secondary forests, clearings, forest margins and shrubby meadows. South of the boundary of continuous zonal broad-leaved forests, this vector is confined to moister floodplain forests and shrubwoods located in depressions of the terrain (Filippova, 1977). The tick I. persulcatus typically inhabits variants of taiga (boreal coniferous forests) but occurs in some other plant formations as well. In the plains its range extends over the area with at least six types of zonal vegetation: northern, midand southern taiga forests, mixed and broad-leaved forests, and forest-steppe. The altitudinal zonality of vegetation in the mountain regions contributes to this diversity. The list of plant formations in which this tick occurs is extensive (Korenberg and Kovalevskii, 1985). In the plain mid-taiga forests, where humidity is considerable and heat supply (seasonal sum of daily positive temperatures) is low, I. persulcatus occurs only in the warmest and most well-drained areas. In the Russian northwest, for example, natural ITBB foci in such a situation are located in spruce and small-leaved forests isolated from each other by numerous lakes, bogs and wet and over-mature stands (Fig. 7.4, 1); in vast plain watershed mid-taiga forests, I. persulcatus occurs mainly in light forest margins and in the immediate vicinity of villages and towns (Fig. 7.4, 2); and in the mid-taiga zone of West Siberia, intrazonal conditions of large river valleys are most favourable for borrelial circulation (Fig. 7.4, 3). Fig. 7.3. (Opposite) Incidence of different Borrelia genospecies and their mixtures in small mammals and ticks. The histograms (1–7) show the portion (in %) of mixed isolates (A), Borrelia afzelii (B), Borrelia garinii 20047T (C) and B. garinii NT29 (D) of the number (indicated in square brackets) of isolates, obtained from: (1) total [595]; (2) Clethrionomys glareolus [132]; (3) small mammals of other species [58]; (4) fed and partly engorged Ixodes persulcatus larvae and nymphs, collected from mammals [85]; (5) unfed I. persulcatus nymphs, collected from vegetation [95]; (6) unfed adult I. persulcatus ticks, collected from vegetation [165]; (7) fed and partly engorged Ixodes trianguliceps ticks, collected from mammals [60].

190

E.I. Korenberg et al.

Fig. 7.4. Scheme of landscape preference of populations of Ixodes persulcatus ticks – plain territories (according to Korenberg and Kovalevskii, 1985). (1) North West region, middle taiga at the south of Karelia; (2) Northern region, middle taiga at the south of Komi Republic; (3) West Siberia, middle taiga at the right bank of the Ob river; (4) Volga–Vyatka region, European southern taiga forests; (5) West Siberia, southern taiga near to the Irtysh river; (6) West Siberia, northern forest–steppe; (7) Ural region, Kungur forest–steppe. Bold black lines under the landscape profiles show the places, where ticks are found.

Ecology of B. burgdorferi in Russia

191

In the southern taiga forests of the European plain, where heat supply is high, I. persulcatus inhabits virtually all plain watershed forest biocenoses (Fig. 7.4, 4). However, when moisture is excessive, as in many areas of the southern taiga subzone of West Siberia, this vector occurs mainly in well-drained river valleys and mixed waterless valley forests (Fig. 7.4, 5). In the forest–steppe, under conditions of some moisture deficiency but sufficient temperatures for I. persulcatus, the tick is only found in the central parts of birch–aspen forest islands (Fig. 7.4, 6) or on the bottoms of ravines and slopes of northern exposure (Fig. 7.4, 7). The patterns of the altitudinal distribution of I. persulcatus ticks and ITBB foci vary considerably depending on such factors as the general geographical location of the mountain system, its orography and altitudinal gradient. In the mountain forests in the south of central Siberia, for example, parasitological prerequisites for the persistence of B. burgdorferi s.l. exist virtually everywhere, from the piedmont forest–steppes to the upper boundary of the forest belt, which is located approximately 1400 m above sea level (Fig. 7.5, 1). A similar picture is observed in the northeastern Altai mountains, but in this case the upper forest boundary is located higher and the taiga tick occurs at elevations of up to 2000 m. However, it does not occur in forest areas that, due to some orographic factors, begin at elevations above 1500 m; the same applies to woodless steppe-

Fig. 7.5. Scheme of landscape preference of populations of Ixodes persulcatus ticks – mountain territories (according to Korenberg and Kovalevskii, 1985). (1) Mountains of the south of Central Siberia; (2) Central Altai; (3) mountains of the north surroundings of the Amur river (Far East); (4) Tien Shan. Bold black lines under the landscape profiles show the places where ticks are found. Vertical scale, height above sea-level.

192

E.I. Korenberg et al.

like slopes of southern, southeastern and southwestern exposure (Fig. 7.5, 2). Near the northern boundary of the I. persulcatus range, in the mountain ridges located in the middle taiga subzone of Khabarovsk territory, this vector regularly occurs on the northwestern slopes at elevations of no more than 750–800 m, whereas on the opposite warmer slopes it is found up to the mountain tops at elevations of approximately 1400 m above sea level (Fig. 7.5, 3). These examples do not exhaust the entire diversity of landscape–biotic preferences of the vector that determine the distribution of ITBB biotopes over the vast and varied territory of Russia. Due to the many factors involved, essentially different conditions may form within a zone, which may be more varied than conditions between zones.

Ecological Risk Factors In diseases with natural focality, the number of persons infected in a focus and the actual risk of infection are eventually the product of the epizootic potential of the focus and the frequency of human contact with it (Korenberg and Yurkova, 1983). In essence, the ecological factors responsible for the risk of infection are the same factors (parameters) that determine the epizootic potential of a focus. In ITBB, as in TBE (Korenberg and Kovalevskii, 1999), the principal parameter is the number of infected ticks, and this number depends on the general level of tick abundance and the prevalence of borreliae in them. The latter parameters differ considerably from one landscape subzone to another, and this is manifested especially clearly in many Russian oblasts that extend from the north to the south for a great distance. Thus, in different subzones of the Tyumen oblast in West Siberia, the average abundance of I. persulcatus adults and the prevalence of borreliae in them differ by factors of 9 and c. 2, respectively (Table 7.6). Differences between these ecological parameters in different landscape subzones of the Perm oblast in the Ural region are of similar magnitude. The frequency of visits by people to the forest is approximately the same throughout this oblast (Table 7.7), but data on the frequency of tick attachment and the parameters of Table 7.6. Parameters of abundance of adult I. persulcatus ticks on vegetation and prevalence of Borrelia burgdorferi sensu lato in them (by dark-field microscopy of internal organs) in different landscape subzones of Tyumen oblast, situated in the West Siberian region. (After Kolchanova, 1997.) Number of ticks per flag-hour Landscape subzones

Minimum Maximum Average

Middle taiga Southern taiga Subtaiga Northern forest–steppe

0 0 1 0

24 37 57 1

1.6 2.7 4.6 0.5

Prevalence of Borrelia spp. in ticks 27.5 32.8 44.6 22.6

± ± ± ±

4.4 3.6 3.4 6.2

Landscape subzones Middle taiga Southern taiga Coniferous–deciduous forests Forest–steppe Mountain taiga

% of people who visited forest 90.8 88.1 84.0 88.1 88.6

± ± ± ± ±

3.7 4.9 4.2 3.8 3.3

% of people who noted tick attachment 17.2 41.3 39.3 38.1 24.9

± ± ± ± ±

4.8 2.9 5.6 5.7 4.5

% of sera with titre 1:20 or more

Mean geometrical titre of antibodies for positive sera

Average long-term parameter of morbidity per 100,000 people

9.2 ± 3.7 27.2 ± 2.6 22.9 ± 4.6 26.2 ± 5.2 12.7 ± 3.5

1:12.1 1:24.2 1:22.6 1:19.7 1:17.1

4.84 16.27 13.51 8.99 11.32

Ecology of B. burgdorferi in Russia

Table 7.7. Frequency of human contacts with ticks and agents of ITBB in different landscape–geographical subdivisions of Perm oblast, situated in the Ural region. (After Alypova et al., 2002.)

193

194

E.I. Korenberg et al.

Table 7.8. Strength of correlations between long-term parameters of adult Ixodes persulcatus density, infected tick density, Borrelia burgdorferi sensu lato prevalence in these ticks and parameters of human ITBB morbidity (observations in the Ural region). Density of adult I. persulcatus (per hectare) Total

Infected ticks only

1993 835 424 50.8 1994 663 243 36.6 1995 226 78 34.5 1996 1423 865 60.8 1997 212 67 31.6 1998 390 115 29.5 1999 1135 546 48.1 2000 463 165 35.7  Coefficient of correlation (r) with parameters of morbidity   0.79 0.85 0.90

Number of ITBB cases (per 100,000 people) 32.3 16.5 15.4 35.7 14.6 17.1 20.8 11.1

  

Year

Borrelia prevalence in ticks (%)

anti-ITBB immunity among residents indicate that people living in the subzones of southern taiga forests, coniferous/broad-leaved forests, and forest–steppe are at a higher risk of acquiring ITBB. Risk is moderate in the southern part of the mountain taiga subzone and is relatively low in the southern part of the middle taiga subzone. There is a close correlation in different landscape subzones between the proportion of people bitten by ticks and the proportion of sera with a high IFA titre of antibodies (> 1:20 and higher) (correlation coefficient r = 0.98) and also between long-term average parameters of ITBB morbidity (r = 0.8) (Alypova et al., 2002). A long-term study in the Ural region revealed that the level of ITBB morbidity and, therefore, the risk for people to acquire infection changed significantly from year to year (Table 7.8). These changes correlate well with annual fluctuations in the total density of adult ticks and the density of ticks infected by Borrelia spp. In contrast to the situation with TBE (Korenberg and Kovalevskii, 1999), there is also a close correlation between ITBB morbidity and Borrelia spp. prevalence in ticks. This may be explained by differences in the infective doses (the dose necessary for causing a clinically manifest disease) of TBE virus and borreliae and in specific features of their behaviour at the earliest stage of the infection process. Nevertheless, the density of infected ticks clearly plays the key role as the principal ecological risk factor.

Conclusion This chapter presents an overview of the available data on the ecology of B. burgdorferi s.l. in Russia. The vast Russian territory, with its diverse landscape,

Ecology of B. burgdorferi in Russia

195

climatic conditions and ecosystems, has been poorly studied in this respect. However, it has already become apparent that the ecology of widespread Borrelia genospecies has very important regional features that will be the focus of research attention in the future. In this connection, we attach special significance to studies on the specificity of the aetiological structure of natural ITBB foci in different regions, the degree of their stability and the factors determining their dynamics.

Acknowledgements This study was supported in part by the Russian Foundation for Basic Research, grant no. 01-04-48152.

References Alekseev, A.N. (1993) Tick–Tick-Borne Pathogen System and its Emergent Qualities. St Petersburg, 203 pp. (in Russian). Alypova, I.I., Korenberg, E.I. and Vorob’eva, N.N. (2002) Risk of ixodid tick-borne borreliosis infection for people in different landscape subzones of Perm region. Meditsinskaya Parasitologiya i Parasitarnye Bolezni 1, 37–40 (in Russian). Apostolova, I.Y., Borisova, T.I. and Kulikova, E.V. (1996) Infection rate of ticks with Borrelia burgdorferi in suburban zone of Irkutsk city. In: Human Diseases with Natural Focality. Omsk, pp. 190–191 (in Russian). Assous, M., Postic, D., Paul, G., Nevot, P. and Baranton, G. (1994) Individualisation of two new genomic groups among American Borrelia burgdorferi sensu lato strains. FEMS Microbiology Letters 121, 93–98. Baranton, G., Postic, D., Saint Girons, I., Boerlin, P., Piffaretti, J.-C., Assous, M. and Grimont, P.A.D. (1992) Delineation of Borrelia burgdorferi sensu stricto, Borrelia garinii sp. nov., and group VS461 associated with Lyme borreliosis. International Journal of Systematic Bacteriology 42, 378–383. Belfaiza, J., Postic, D., Bellenger, E.,

Baranton, G. and Saint Girons, I. (1993) Genomic fingerprinting of Borrelia burgdorferi sensu lato by pulsedfield gel electrophoresis. Journal of Clinical Microbiology 31, 2873–2877. Belyaeva, M.L. (1998) Results of studies on Borrelia prevalence in ticks in Sverdlovsk region. In: Actual Problems of Infections with Natural Focality. Izhevsk, pp. 166–167 (in Russian). Burenkova, L.A. (2000) Seasonal changes in infection rate of ticks with agents of tick-borne borrelioses at the north of Kaluga region. Dezinfektsionnoe Delo 2, 14–16 (in Russian). Burgdorfer, W., Barbour, A., Hayes, S., Peter, O. and Aeschlimann, A. (1983) Erythema chronicum migrans – a tickborne spirochetosis. Acta Tropica 40, 79–83. Dubinina, E.V. and Alekseev, A.N. (1999) The biodiversity dynamics of the pathogens being transmitted by the ticks of the genus Ixodes: an analysis of the long-term data. Meditsinskaya Parasitologiya i Parasitarnye Bolezni 2, 13–19 (in Russian). Filippova, N.A. (1977) Ixodid Ticks of Subfamily Ixodinae: Arachnaidae. (Fauna of USSR. Vol. 4, iss. 4.) Nauka, Moscow, Leningrad, 396 pp. (in Russian).

196

E.I. Korenberg et al.

Filippova, N.A. (1999) Sympatry of closely related species of Ixodid ticks and its possible role in parasitic systems of natural foci of transmissive diseases. Parazitologiya 33, 223–241 (in Russian). Fukunaga, M., Sohnaka, M. and Yanagihara, Y. (1993) Analysis of Borrelia species associated with Lyme disease by rRNA gene restriction fragment length polymorphism. Journal of General Microbiology 139, 1141–1146. Gern, L., Humair, P.F., Hu, C.M. and Leuba-Garcia, S. (1997) Ecology of Borrelia burgdorferi sensu lato in Europe. In: Suss, J. and Kahl, O. (eds) Tick-borne Encephalitis and Lyme Borreliosis. Pabst Science Publishers, Lengerich, pp. 271–280. Gorelova, N.B. (1993) Museum of Borrelia of the Russian Center for Borrelioses. In: Korenberg, E.I. (ed.) Problems of Tickborne Borrelioses. Moscow, pp. 38–44. Gorelova, N.B., Korenberg, E.I., Kovalevskii, Y.V. and Shcherbakov, S.V. (1995) Small mammals as reservoir hosts for Borrelia in Russia. Zentralblatt für Bacteriologie 282, 315–322. Gorelova, N.B., Korenberg, E.I., Kovalevskii, Y.V., Postic, D. and Baranton, G. (1996) Isolation of Borrelia from the tick Ixodes trianguliceps (Ixodinae) and the significance of this species in epizootiology of ixodid-tick-borne borrelioses. Parazitologiya 30, 13–18. Gorelova, N.B., Korenberg, E.I., Filippova, N.A. and Postic, D. (2001a) First isolation of Borrelia, pathogenic for human, from Ixodes pavlovskyi Pom. ticks. Doklady Rossiyskoyi Akademii Nauk 378, N 4, 1–2 (in Russian). Gorelova, N.B., Korenberg, E.I., Postic, D., Yunicheva, Y.V. and Ryabova, T.E. (2001b) First isolation of Borrelia burgdorferi sensu stricto in Russia. Zhurnal Microbiologii, Epidemiologii i Immunobiologii 3, 459–462 (in Russian). Gorin, O.Z., Zlobin, V.I., Arbatskaya, E.V., Dzhioev, Y.P., Kozlova, I.V., Ivanova, L.V., Chernogor, L.I.,

Cuntsova, O.V. and Verhozina, M.V. (1996) Actual problems of epidemiology and prophylaxis of transmissible infections in Siberia. Zhurnal Infektsionnoy Patologii 3, 10–12 (in Russian). Grigoryeva, L.A. (1996) Shrews as reservoir hosts for Borrelia in the north-west of Russia. Parazitologiya 30, 458–459 (in Russian). Grigoryeva, L.A. and Tretyakov, K.A. (1998) Peculiarity of parasitic system ixodid ticks – Borrelia – micromammalia in the north-west of Russia. Parazitologiya 32, 422–430 (in Russian). Hu, C.M., Humair, P.F., Wallich, R. and Gern, L. (1997) Apodemus sp. rodents, reservoir hosts for Borrelia afzelii in an endemic area in Switzerland. Zentralblatt für Bacteriologie 285, 558–564. Humair, P.F., Peter, O., Wallich, R. and Gern, L. (1995) Strain variation of Lyme disease spirochetes isolated from Ixodes ricinus ticks and rodents collected in two endemic areas in Switzerland. Journal of Medical Entomology 32, 433–438. Humair, P.F., Rais, O. and Gern, L. (1999) Transmission of Borrelia afzelii from Apodemus mice and Clethrionomys voles to Ixodes ricinus ticks: differential transmission pattern and overwintering maintenance. Parasitology 118, 33–42. Ivanov, L.I., Liberova, R.N., Korenberg, E.I., Zdanovskaya, N.I., Vorob’eva, R.N., Volkov, V.I. and Vysochina, N.P. (1996) Natural foci of Lyme disease in Khabarovsk region. In: International Scientific Conference ‘Tick-borne Viral, Rickettsial and Bacterial Infections’. Irkutsk, pp. 89–90. Johnson, R.C., Schmid, G.P., Hyde, F.W., Steigerwalt, A.G., and Brenner, D.J. (1984) Borrelia burgdorferi sp. nov.: etiologic agent of Lyme disease. International Journal of Systematic Bacteriology 34, 496–497. Kolchanova, L.P. (1997) Spontaneous Borrelia infection in ticks and degree of their individual infectivity in different

Ecology of B. burgdorferi in Russia

landscape subzones of Tyumen region. Meditsinskaya Parasitologiya i Parasitarnye Bolezni 1, 49–50 (in Russian). Korenberg, E.I. (1962) Role of birds in feeding of ixodid ticks in natural foci of encephalitis in forest zone. Zoologicheskii Zhurnal 41, 1220–1225 (in Russian). Korenberg, E.I. (1964) Birds and the problem of natural focality of tick-borne encephalitis. Zoologicheskii Zhurnal 45, 245–260 (in Russian). Korenberg, E.I. (1978) Mammals – hosts of adult taiga ticks at the east of Russian plain. In: Panteleev, P.A. (ed.) Proceedings of the 2nd Congress of All-Union Teriological Society. Nauka, Moscow, pp. 257–258 (in Russian). Korenberg, E.I. (1979) Small mammals and problem of natural focus of tickborne encephalitis. Zoologicheskii Zhurnal 58, 542–552 (in Russian). Korenberg, E.I. (1993) Borrelioses. In: Pokrovsky, V. (ed.) Manual on Epidemiology of Infectious Diseases, Vol. 2. Meditsina, Moscow, pp. 382–392 (in Russian). Korenberg, E.I. (1994) Comparative ecology and epidemiology of Lyme disease and tick-borne encephalitis in the former Sovet Union. Parasitology Today 10, 157–160. Korenberg, E.I. (1995) Ixodid tick-borne borrelioses (ITBBs), infections of the Lyme borreliosis group, in Russia: country report. In: Reports of WHO Workshop on Lyme Borreliosis Diagnosis and Surveillance. Sanitati, WHO/CDC/ VPH/95.141-1, Warsaw, pp. 128–136. Korenberg, E.I. (1996) Taxonomy, phylogenetic relations and centres of origin of distinct forms of spirochetes of the genus Borrelia transmitted by ixodid ticks. Uspehi Sovremennoy Biologii 116, 389–406 (in Russian). Korenberg, E.I. (1999) Interaction between transmissible disease agents in ixodid ticks (Ixodidae) with a mixed infection. Parazitologiya 33, 273–289 (in Russian). Korenberg, E.I. and Kovalevskii, Y.V. (1981) Regionalization of the Tick-borne

197

Encephalitis Range. Advances in Science and Technology, Medical Geography, Vol. 11. VINITI, Moscow, 148 pp. Korenberg, E.I. and Kovalevskii, Y.V. (1985) Landscape relation. In: Filippova, N.A. (ed.) Taiga Tick Ixodes persulcatus Schulze (Acarina, Ixodidae). Nauka, Leningrad, pp. 193–198 (in Russian). Korenberg, E.I. and Kovalevskii, Y.V. (1994) A model for relationships among the tick-borne encephalitis virus, its main vectors, and hosts. In: Harris, K.E. (ed.) Advances in Disease Vector Research 10. Springer-Verlag, New York, pp. 65–92. Korenberg, E.I. and Kovalevskii, Y.V. (1999) Main features of tick-borne encephalitis eco-epidemiology in Russia. Zentralblatt für Bacteriologie 289, 525–539. Korenberg, E.I. and Kryuchechnicov, V.N. (1996) Ixodid tick-borne borrelioses – a new group of human diseases. Zhurnal Microbiologii, Epidemiologii i Immunobiologii 4, 104–108 (in Russian). Korenberg, E.I. and Lebedeva, N.N. (1969) Distribution and some general features of the ecology of Ixodes trianguliceps Bir. in the Soviet Union. Folia Parasitologica 16, 143–152. Korenberg, E.I. and Yurkova, E.V. (1983) The problem of predicting epidemic manifestations of natural foci of human diseases. Meditsinskaya Parasitologiya i Parasitarnye Bolezni 3, 3–10 (in Russian). Korenberg, E.I., Kryuchechnikov, V.N., Dekonenko, E.P., Scherbakov, S.V. and Anan’ina, Y.V. (1986) Serological verification of Lyme disease in Russia. Zhurnal Microbiologii, Epidemiologii i Immunobiologii 6, 111–113 (in Russian). Korenberg, E.I., Kryuchechnikov, V.N., Kovalevskii, Y.V., Scherbakov, S.V., Kuznetsova, R.I. and Levin, M.L. (1987a) Tick Ixodes persulcatus Schulze – a new vector of Borrelia burgdorferi. Doklady Academii Nauk SSSR 297, 1268–1270 (in Russian).

198

E.I. Korenberg et al.

Korenberg, E.I., Scherbakov, S.V. and Kryuchechnikov, V.N. (1987b) Data on distribution of Lyme disease in USSR. Meditsinskaya Parasitologiya i Parasitarnye Bolezni 2, 71–73 (in Russian). Korenberg, E.I., Kovalevskii, Y.V., Kuznetsova, R.I., Fonarev, L.S., Churilova, A.A., Antykova, L.P., Kalinin, M.I., Kryuchechnikov, V.N., Mebel’, B.D., Scherbakov, S.V. and Kovtunenko, S.S. (1988) Revealing and first results of studies of Lyme disease at the north-west of USSR. Meditsinskaya Parasitologiya i Parasitarnye Bolezni 1, 45–48 (in Russian). Korenberg, E.I., Scherbakov, S.V., Zaharycheva, T.A., Levin, M.L., Kalinin, M.I. and Kryuchechnikov, V.N. (1989) Lyme disease in Khabarovsk region. Meditsinskaya Parasitologiya i Parasitarnye Bolezni 5, 74–78 (in Russian). Korenberg, E.I., Kovalevskii,Y.V., Kryuchechnikov, V.N. and Gorelova, N.B. (1991a) Tick Ixodes persulcatus Schulze, 1930 as a vector of Borrelia burgdorferi. In: Dusbabek, F. and Bukva, V. (eds) Modern Acarology, Vol. 1. Academia, Prague, pp. 119–123. Korenberg, E.I., Kuznetsova, R.I., Kovalevskii, Y.V., Vasilenko, Z.E. and Mebel’, B.D. (1991b) Main features of epidemiology of Lyme disease at the north-west of USSR. Meditsinskaya Parasitologiya i Parasitarnye Bolezni 3, 14–17 (in Russian). Korenberg, E.I., Gorelova, N.B., Postic, D., Kovalevskii, Y.V., Baranton, G. and Vorobyeva, N.N. (1997) Reservoir hosts and vectors of Borrelia, causative agents of ixodid tick-borne borrelioses in Russia. Zhurnal Microbiologii, Epidemiologii i Immunobiologii 6, 36–38 (in Russian). Korenberg, E.I., Gorelova, N.B., Postic, D. and Kotti, B.K. (1999) Borrelia species, new for Russia, as possible causative agents of ixodid tick-borne borreliosis. Zhurnal Microbiologii, Epidemiologii i Immunobiologii 2, 3–5 (in Russian).

Korenberg, E.I., Kovalevskii, Y.V., Levin, M.L. and Shchyogoleva, T.V. (2001a) The prevalence of Borrelia burgdorferi sensu lato in Ixodes persulcatus and I. ricinus ticks in the zone of their sympatry. Folia Parasitologica 48, 63–68. Korenberg, E.I., Gorban’, L.J., Kovalevskii, Y.V., Frizen,V.I. and Karavanov, A.S. (2001b) Risk of human tick-borne encephalitis, borrelioses, and double infection in the Pre-Ural Region of Russia. Emerging Infectious Diseases (in press). Kryuchechnikov, V.N., Korenberg, E.I., Scherbakov, S.V., Kovalevskii, Y.V. and Levin, M.L. (1988) Identification of Borrelia, isolated in USSR from Ixodes persulcatus Schulze ticks. Zhurnal Microbiologii, Epidemiologii i Immunobiologii 12, 41–44 (in Russian). Kryuchechnikov, V.N., Korenberg, E.I., Sherbakov, S.V., Gorelova, N.B., Jurzicova, Z., Halouzka, Z., Hubalek, Z., Kovalevskii, Y.V., Levin M.L. and Bunikis, A.I. (1990) Identification of Borrelia, isolated in USSR and Czechoslovakia from Ixodes ricinus (L.) ticks. Zhurnal Microbiologii, Epidemiologii i Immunobiologii 6, 10–13 (in Russian). Kryuchechnikov, V.N., Gorelova, N.B. and Sherbakov, S.V. (1993) Identification of Borrelia and results of studying isolates of Lyme disease agent from Russia and neighbouring countries. In: Korenberg, E.I. (ed.) Problems of Tickborne Borrelioses. Moscow, pp. 45–55 (in Russian). Kurtenbach, K., Peacey, M., Rijpkema, S.G.T., Hoodless, A.N., Nuttall, P.A. and Randolph, S.E. (1998) Differential transmission of the genospecies of Borrelia burgdorferi sensu lato by game birds and small rodents in England. Applied and Environmental Microbiology 64, 1169–1174. Lisovskaya, N.D. (1958) Data on studies of chronical atrophic acrodermatitis. Candidate Sc. thesis, State Institute for Perfection of Physicians, Leningrad, USSR (in Russian).

Ecology of B. burgdorferi in Russia

Liveris, D., Gazumyan, A. and Schwartz, I. (1995) Molecular typing of Borrelia burgdorferi sensu lato by PCR– restriction fragment length polymorphism analysis. Journal of Clinical Microbiology 33, 589–595. Masuzawa, T., Iwaki, A., Sato, Y., Miyamoto, K., Korenberg, E.I. and Yanagihara, Y. (1997) Genetic diversity of Borrelia burgdorferi sensu lato isolated in Far Eastern Russia. Microbiology and Immunology 41, 595–600. Matuschenko, A.A., Venediktov, V.S., Yakimenko, V.V. and Tantsev, A.K. (1996) Role of small mammals in circulation of Borrelia in natural foci of tickborne borreliosis at the south-west of Siberia. Zhurnal Infektsionnoy Patologii 3, 37–39 (in Russian). Mescherskii, G.I. (1898) To the problem of atrophia cutis idiopathica progressiva acuisita. Records of Moscow Venereologic and Dermatologic Society 8, 135–136 (in Russian). Mishin, A.V. (1964) Terms of existence of taiga tick in Udmurtiya ASSR. In: Collected Works of Izhevsk Medical Institute 20. Izhevsk, pp. 66–73 (in Russian). Naumov, R.L. (1985) Trophic relations. In: Filippova, N.A. (ed.) Taiga Tick Ixodes Persulcatus Schulze (Acarina, Ixodidae). Nauka, Leningrad, pp. 277–312 (in Russian). Naumov, R.L. (1999) Tick-borne encephalitis and Lyme disease: epizootiological parallels and monitoring. Meditsinskaya Parasitologiya i Parasitarnye Bolezni 2, 20–26 (in Russian). Naumov, R.L., Gutova, V.P., Ershova, A.S. and Papel’nitskaya, N.P. (1994) Infection rate of taiga ticks with Borrelia in Western Sayans. Meditsinskaya Parasitologiya i Parasitarnye Bolezni 3, 19–20 (in Russian). Nikol’skii, P.V. (1896) Demonstration of the case of atrophia cutis maculata. Diary of the 6th Congress of Russian Physicians in Memory of N.I. Pirogov 12, 89 (in Russian).

199

Pisemskii, N.N. (1902) A case of idiopathic acquired atrophia cutis. Meditsinskoe obozrenie 37, 710–713 (in Russian). Pospelov, A.I. (1905) Manual on Studies of Skin Diseases. Moscow, 650 pp. (in Russian). Pospelov, A.I. (1906) Two cases of autonomous atrophia cutis in adult humans. Meditsinskoe obozrenie 26, 565–574 (in Russian). Postic, D., Assous, M.V., Grimont, P.A.D. and Baranton, G. (1994) Diversity of Borrelia burgdorferi sensu lato evidenced by restriction fragment length polymorphism of rrf (5S) –rrl (23S) intergenic spacer amplicons. International Journal of Systematic Bacteriology 44, 743–752. Postic, D., Korenberg, E., Gorelova, N., Kovalevskii, Y.V., Bellenger, E. and Baranton, G. (1997) Borrelia burgdorferi sensu lato in Russia and neighbouring countries: high incidence of mixed isolates. Research in Microbiology 148, 691–702. Rudakova, S.A. and Matuschenko, A.A. (1996) Results of studies of natural foci of tick-borne borreliosis at regions of the south of Western Siberia. In: Human Diseases with Natural Focality. Omsk, pp. 169–173 (in Russian). Rybakova, N. and Evsyukova, N. (1999) Lyme borreliosis in Vologda province – clinical and epidemiological characteristics. In: VIII International Conference on Lyme Borreliosis and other Emerging Tickborne Diseases, Abstract Book. Munich, p. 43. Saint Girons, I., Gern, L., Gray, J.S., Guy, E.C., Korenberg, E., Nuttall, P.A., Rijpkema, S.G.T., Schönberg, A., Stanek, G. and Postic, D. (1998) Identification of Borrelia burgdorferi sensu lato species in Europe. Zentralblatt für Bacteriologie 287, 190–195. Sato, Y., Miyamoto, K., Iwaki, A., Masuzawa, T., Yanagihara, Y., Korenberg, E., Gorelova, N.B., Volkov, V.I., Ivanov, L.I. and Liberova, R.N. (1996) Prevalence of Lyme disease

200

E.I. Korenberg et al.

spirochetes in Ixodes persulcatus and wild rodents in Far Eastern Russia. Applied and Environmental Microbiology 62, 3387–3389. Savitskii, B.P. (1970) Types of tick-borne encephalitis foci in Khabarovsk region. Scientific Reports of Higher School. Biological Sciences 7, 26–30 (in Russian). Semenov, A.V., Alekseev, A.N. and Dubinina, H.V. (2000) Electrophoretypes, genotypes and Borrelia agents of Ixodes persulcatus ticks. In: International Symposium ‘Ecological Parasitology on the Turn of Millennium’. St Petersburg, p. 56. Smirnova, N.N., Kilevoi, L.Y. and Kushnarev, I.G. (1996) Studies on the state of a natural focus of Lyme disease in Kurgan region. In: Human Diseases with Natural Focality. Omsk, pp. 177–178 (in Russian).

Tarasov, V.N., Ul’yantsenko, E.V., Kondrashova, O.N., Bobrovskaya, G.N. and Oseledchik, N.G. (2000) Prevalence of Borrelia in ixodid ticks in Chelyabinsk region. Meditsinskaya Parasitologiya i Parasitarnye Bolezni 4, 48 (in Russian). Tupikova, N.V., Suvorova, L.G. and Korenberg, E.I. (1980) On the evaluation of the importance of different small mammals species as hosts of larvae and nymphs of Ixodes persulcatus. In: Fauna and Ecology of the Rodents, 14. Moscow University Press, Moscow, pp. 158–176 (in Russian). Vasilieva, I.S. and Naumov, R.L. (1996) Lyme disease parasitosic system, the state of the problem. Communication 1. Pathogens and vectors. Acarina 4, 53–75 (in Russian).

Ecology of Borrelia burgdorferi sensu lato in Japan and East Asia

8

Kenji Miyamoto1 and Toshiyuki Masuzawa2 1

Department of Parasitology, Asahikawa Medical College, Asahikawa, 078-8510, Japan; 2Department of Microbiology, School of Pharmaceutical Sciences, University of Shizuoka, 52-1 Yada, Shizuoka 422-8526, Japan

Introduction The multisystemic disorder known as Lyme disease (Lyme borreliosis, LB) was first described in the USA in the 1970s (Steere et al., 1977), the deer tick (Ixodes ricinus complex) infected with the spirochaete Borrelia burgdorferi sensu lato (s.l.) being the cause of this disease (Steere, 1989). B. burgdorferi s.l. has been classified into at least 11 genetically similar pathogenic species, including B. burgdorferi sensu stricto (s.s.), Borrelia garinii and Borrelia afzelii (Baranton et al., 1992; Canica et al., 1993). While LB is the most frequently reported arthropod-borne infection in North America and Europe (Jaenson, 1991; Centers for Disease Control, 1992; O’Connell et al., 1998; Orloski et al., 2000), the extent of the ecological impact of this disease in Asia has only recently been realized. The study of LB in Japan started in 1987 after an Ixodes persulcatus tick bite in Nagano Prefecture (Kawabata et al., 1987). This was the first recorded case of erythema migrans at the site of the tick bite along with antibodies to B. burgdorferi. However, cases in which there was the appearance of erythema after tick bites from forest regions in Japan, and whether these were allergic or infectious reactions, had been discussed among dermatologists before 1987. According to Yamaguti, 15 tick species (six Ixodes, five Haemaphysalis, one Amblyomma, one Rhipicephalus and two Argas) have been reported from human tick bite cases across Japan (Yamaguti, 1989). Outbreaks of tick bites are observed in all seasons in the southern parts of Japan and from late spring to early autumn in northern parts. The taiga tick I. persulcatus has been incriminated as the potential vector of LB in the Eurasian zone including Japan. However, Ixodes ovatus is the most widely distributed tick in flatlands from the northernmost island of Hokkaido to the southernmost island of Kyushu. In Hokkaido, there are more © 2002 CAB International. Lyme Borreliosis: Biology, Epidemiology and Control (eds J. Gray, O. Kahl, R.S. Lane and G. Stanek)

201

202

K. Miyamoto and T. Masuzawa

areas of natural vegetation than in other areas of Japan; these areas provide the habitat for the birds and mammals that act as the reservoir for the large numbers of ixodid ticks. Recently, the population of sika deer (Cervus nippon) in Hokkaido has increased and has been expanding into new areas with resultant damage to agriculture (Takahashi et al., 1997; Kaji et al., 2000). Numerous cases of tick bites, some accompanied by LB, occur every year, not only in the inhabitants but also in the many visitors to Hokkaido. On 1 April 1999, new legislation for infectious diseases was introduced requiring a clinical doctor to notify any LB cases to the national government. This chapter is devoted to the natural history of the agent, B. burgdorferi s.l., in Japan and East Asia and to the relationships between the spirochaete, its tick vector and its host in nature.

Activity of Ticks and Outbreaks of Tick-bite Cases The activities of host-seeking ticks have been investigated at Furano, central Hokkaido, by blanket-drag sampling over 2.4 km by two people twice a month (Fig. 8.1). Tick questing starts in mid- or late April, depending on weather conditions. Two predominant species of tick, I. persulcatus and I. ovatus, were collected from the study area and the numbers of the ticks increased as the season progressed. The maximum number of ticks was found in early June; thereafter, the number decreased towards the autumn and no ticks were collected in September. The collected ticks were almost all adults but included a small

Fig. 8.1. Seasonal activity of ixodid ticks at Furano from vegetation sampling data (2.4 km) and monthly records of tick-bite cases in Hokkaido.

Ecology of B. burgdorferi in East Asia

203

Table 8.1. Seasonal fluctuation of cases of tick bite in Hokkaido (1980–2000). Month

No. of cases Ixodes persulcatus Ixodes ovatus

Total no.

March April May June July August September October November

0 12 141 225 71 17 4 1 0

0 6 33 29 4 4 2 2 0

0 18 175 (1)a 255 (1)b 75 21 6 3 0

Total

471 (85.2%)

80 (14.5%)

553 (2)

a b

One case of Haemaphysalis flava. One case of Ixodes nipponensis.

number of nymphal I. persulcatus between late April and early July (Miyamoto and Nakao, 1991). All stages of I. persulcatus were collected from vegetation by the flag-sampling method in Niigata Prefecture including Sado Island, and peak numbers of ticks were observed between summer and early autumn (Saito, 1959). The questing activity of I. persulcatus in each locality is quite different. In general, the life cycle of I. persulcatus usually takes 2 years in Japan, but it was completed in only 4.5 months when the tick was reared in the laboratory (Saito, 1960). Since the first tick specimen was obtained from the back of a human subject in 1980, a total of 553 ticks have been collected from cases between each April and October, from 1980 to 2000, with maximum numbers in June in Hokkaido (Table 8.1). Of these ticks, 471 (85.2%) out of 553 specimens were I. persulcatus, and 80 (14.5%) were I. ovatus, the remaining two being Ixodes nipponensis and Haemaphysalis flava (Table 8.1; unpublished data). Almost all cases were infested with female ticks, except for one male and two nymphs of I. persulcatus, and most were found as single specimens, although very rarely two or three occurred together. I. persulcatus prevails on flatland from Hakodate in southern Hokkaido to Wakkanai in the northernmost part of the island, whereas on the mainland this tick species inhabits the high altitude mountainous zone (an elevation of 800–1000 m) from Aomori to Oita. There is a particularly high density of I. persulcatus in the mountainous areas of Nagano Prefecture (Uchikawa et al., 1991).

B. burgdorferi s.l. in Japan and Relationships with Vector Species Since the initial report of LB in Japan (Kawabata et al., 1987), several B. burgdorferi s.l. species have been isolated from patients, wild rodents, migratory birds

204

K. Miyamoto and T. Masuzawa

and five species of ixodid ticks (I. persulcatus, I. ovatus, Ixodes tanuki, Ixodes turdus and Ixodes columnae) (Masuzawa et al., 1991a,b, 1994; Miyamoto et al., 1991, 1992a,b, 1997, 2000; Nakao and Miyamoto, 1993a,b,c, 1994; Miyamoto and Nakao, 1994; Hashimoto et al., 1995). Recently developments in molecular biology have provided new and powerful epidemiological tools for the accurate typing of B. burgdorferi s.l. Serotyping systems based on reactivity with monoclonal antibodies (mAbs), various DNA hybridization methods, as well as restriction fragment length polymorphism (RFLP) analysis using several spirochaete rRNA genes, rRNA gene intergenic spacers, the flagellin gene and genomic DNA as targets, have been successfully exploited (Baranton et al., 1992; Marconi and Garon, 1992; Wilske et al., 1992; Fukunaga et al., 1993b; Kawabata et al., 1994; Postic et al., 1994; Fukunaga and Koreki, 1996). Using these powerful methods, we have begun to systematically explore the distribution of many B. burgdorferi s.l. isolates in the hope that such information will aid our understanding of the Borrelia spp. that cause LB in Japan. Epidemiological surveys have revealed the high prevalence of LB-related Borrelia spp. in both I. persulcatus and I. ovatus, which are the most common tick species that infest humans in Japan (Uchikawa et al., 1991; Miyamoto et al., 1992a,b; Nakao et al., 1992a,b). Divergent protein profiles for several Japanese B. burgdorferi s.l. isolates from I. persulcatus have been found, some of which share similar protein profile patterns with some European isolates (Masuzawa et al., 1991b; Nakao et al., 1992a; Fukunaga et al., 1993a). RFLP patterns of the 16S rRNA gene of isolates from I. persulcatus can be classified into five groups (ribotypes), II–VI, whereas RFLP analysis of B. burgdorferi s.s., B. garinii and B. afzelii from North America and Europe reveals three ribotypes, I, II and III. While some of the Japanese isolates from I. persulcatus belonging to ribotypes II and III were identified as B. garinii and B. afzelii, the species identity of the other isolates generating the ribotype IV, V and VI RFLP patterns was not clear (Fukunaga et al., 1993a,b; Nakao et al., 1994a,b). However, because these isolates were subsequently found to have 82–85% DNA homology with B. garinii type strain 20047 and over 98% similarity with the fla gene of B. garinii strain Ip90, we place these isolates (ribotypes IV, V and VI), which have not been found in America and Europe, as members of B. garinii (Fukunaga and Koreki, 1996; Masuzawa et al., 1998). However, the picture is not yet clear-cut as genetic and immunological heterogeneity have been reported for many European B. garinii isolates (Wilske et al., 1993; Bunikis et al., 1995; Will et al., 1995). At least 61 strains of B. burgdorferi s.l. have been classified in Japan on the basis of reactivity with 16 mAbs to outer surface protein A (OspA) (Masuzawa et al., 1996b). Whereas all isolates of Japanese B. afzelii examined thus far can be classified into one OspA serotype, at least nine OspA serotypes have been identified as B. garinii. It has been reported by Wilske et al. (1993) that European B. garinii can be classified into at least five Osp serotypes. However, none of the Japanese strains identified as B. garinii has been found to have identical reactivity with the European B. garinii strains. Using a specific panel of B. garinii mAbs designed for this purpose, Will et al.

Ecology of B. burgdorferi in East Asia

205

(1995) clearly established the immunological differences between the Japanese and European B. garinii isolates. These findings are also supported by DNA sequencing studies of the OspA genes of the spirochaetes (T. Masuzawa, unpublished data). On the other hand, all isolates identified as B. afzelii in Japan were homologous (and were also similar to European B. afzelii), showing uniformity in the molecular sizes of OspA and OspB and in mAb reactivity. Many B. burgdorferi s.l. isolates worldwide have been compared, characterized and identified by analysing the RFLP patterns of the 5S–23S rRNA gene intergenic spacer, a method originally pioneered by Postic et al. (1994). Based on the RFLP 5S–23S spacer patterns, we have compiled data summarizing the classification of representative B. burgdorferi s.l. isolates from Japan and other areas in the world (Table 8.2). To our knowledge, most European B. garinii generated RFLP pattern B (by digestion of MseI). Furthermore, B. garinii from Japan (Masuzawa et al., 1996a), Far Eastern Russia (Masuzawa et al., 1997a; Postic et al., 1997), the western part of Russia (Postic et al., 1997), China (Li et al., 1998) and Korea (Masuzawa et al., 1999) showed not only pattern B or its variation (Bvl, Bv2), but also pattern C or its variation (Cvl), which Postic et al. (1997) referred to as B. garinii variant NT29, the first isolate of which originated from I. persulcatus collected in the Nagano prefecture, Japan. In addition, a few isolates from China, Korea and Japan (Ishiguro et al., 2000) displayed pattern R and Rv1. More importantly, isolates generating pattern C and R have not been found in any European isolates from I. ricinus (Postic et al., 1997; T. Masuzawa et al., unpublished data). It is of interest that pattern C (B. garinii variant NT29) was the most common B. garinii isolate from I. persulcatus in Japan as it was in most of those isolates classified in the ribotype IV group. Therefore, based on the RFLP analysis data reported here, we have tentatively classified Japanese B. garinii into two subtypes. These are the Eurasia-type B. garinii subtype (type strain 20047), which is found in Europe and Asia and generates the RFLP pattern B, and the Asia-type B. garinii subtype (variant strain NT29) found only in Asia, which generates pattern C by RFLP. In Japan, the Asia-type B. garinii is not only the most commonly found Borrelia spp., but also the most frequent spirochaete isolated from patients. I. ovatus, which belongs to the subgenus Partipalpiger (Hoogstraal et al., 1973), is another common and widely distributed ixodid tick that infests human habitats in Japan. B. burgdorferi s.l. isolated from I. ovatus shows homology in the molecular size of OspA and OspB and by reactivity to mAb H5332 specific to the B. burgdorferi species complex (Nakao et al., 1992a; Masuzawa et al., 1994). Although the classification of relapsing fever Borrelia spp. has been based on vector tick species, to our knowledge, the specific relationship between LB-related Borrelia spp. and tick species has not yet been fully characterized. Based on an extensive study comparing several B. burgdorferi s.l. isolates from I. ovatus and a variety of reservoir animal species (Masuzawa et al., 1995), it appears that the isolates were homologous in protein profiles and OspC gene sequences (Fukunaga and Takahashi, 1994; Masuzawa et al., 1997b). However, these isolates were different from the other LB Borrelia spp. such as B. burgdorferi s.s., B. garinii and B.

Species B. burgdorferi s.s. B. garinii

B. japonica B. valaisiana B. valaisiana-related

B31 20047 NP81 HT59 NT31 XB190 XB141 NT29 ASF KKU5 XB37 ChY13p 935T VS461 KKU1 17Y NT28 HO14 VS116 Am501 5MT KR3, KR6 CKA2b

Vector tick

Isolated in:

I. scapularis I. ricinus I. persulcatus I. persulcatus I. persulcatus I. persulcatus I. persulcatus I. persulcatus I. persulcatus I. nipponensis I. persulcatus I. persulcatus I. persulcatus I. ricinus I. nipponensis I. granulatus I. persulcatus I. ovatus I. ricinus I. columnae I. nipponensis R. losea A. agrarius

USA Europe Japan Japan Japan Far eastern Russia Far eastern Russia Japan Japan Korea Far eastern Russia Northern China Korea Europe Korea Korea Japan Japan Europe Japan Korea Taiwan China

254 253 255 255 255 250 251 253 253 253 253 237 237 246 246 248 246 236 255 249 254 254 254

A′ B′ B′ NS NS B′v1 B′v2 C′ C′ C′ C′v1 R′ R′ D′ D′ D′v1 N′ NS B″ NS S′ S′ S′

DraI restriction fragment sizes (bp) 144, 53, 29, 28 201, 52 203, 52 203, 52 203, 52 201, 49 201, 50 144, 57, 52 144, 57, 52 144, 57, 52 157, 52, 44 185, 52 185, 52 174, 52, 20 174, 52, 20 175, 52, 21 144, 52, 30, 20 No fragment 203, 52 No fragment 173, 81 173, 81 173, 81

MseI MseI RFLP restriction fragment pattern sizes (bp) A B B B B Bv1 Bv2 C C C Cv1 R Rv1 D D Dv1 N E F Q S S S

108, 108, 108, 108, 108, 107, 107, 108, 107, 107, 107, 105, 107, 108, 108, 108, 107, 108, 175, 169, 107, 107, 107,

51, 95, 96, 96, 96, 95, 95, 57, 57, 57, 51, 79, 79, 68, 68, 68, 51, 78, 50, 51, 58, 58, 58,

38, 50 51 51 51 48 49 50, 51, 51, 51, 53 51 50, 50, 51, 38, 50 23, 23, 43, 43, 43,

29, 28

38 38 38 44

20 20 21 30, 20 7 6 24, 22 24, 22 24, 22

K. Miyamoto and T. Masuzawa

B. afzelii

Strain

Amplicon size (bp)

DraI RFLP pattern

206

Table 8.2. Borrelia burgdorferi sensu lato identified by RFLP analysis of the 5S–23S intergenic rRNA gene spacer amplicons.

10MT CKA2a TA1, TA2

B. andersonii B. lusitaniae B. bissettii

B. sinica

Korea China Taiwan

254 254 254

S′v1 S′v1 S′v1

173, 81 173, 81 173, 81

Sv1 Sv1 Sv1

150, 58, 24, 22 150, 58, 24, 22 150, 58, 24, 22

CKG5 TM1

I. granulatus M. formosanus

China Taiwan

254 254

S′v1 S′v1

173, 81 173, 81

Sv1 Sv1

150, 58, 24, 22 150, 58, 24, 22

CKA3a KR1, KR2, KR4 TM2 CKA1 CKG4 Hk501 Ya501

A. agrarius R. losea M. formosanus A. agrarius I. granulatus I. tanuki I. turdus

China Taiwan Taiwan China China Japan Japan

246 246 246 253 253 245 248

S′v2 S′v2 S′v2 S′v3 S′v3 O′ P′

144, 144, 144, 144, 144, 173, 144,

Sv2 Sv2 Sv2 Sv3 Sv3 O P

I. I. I. I. I. I.

dentatus ricinus ricinus neotomae neotomae scapularis

USA Europe Europe USA USA USA

266 257 255 257 226 253

NS G′ H′ I′ J′ K′

No fragment 145, 83, 29 158, 81, 16 144, 53, 33, 27 144, 53, 29 173, 53, 27

N. confucianus N. confucianus I. ovatus

China China China

235 236 236

145, 59, 28, 145, 59, 28, 145, 59, 28, 145, 51, 29, 145, 51, 29, 174, 51, 20 107, 51, 38, 16, 8, 7 120, 67, 51, 108, 81, 39, 108, 79, 52, 108, 51, 38, 108, 51, 38, 108, 51, 34, 17, 12, 4 107, 48, 38, 107, 49, 38, 107, 49, 38,

21123 PotiB1, PotiB2 PotiB3 DN127 California strains 25015 CMN3 CMN1a, CMN2 CWO1

102 102 102 52, 29, 28 52, 29, 28 52, 20 81, 23

157, 49, 29 144, 50, 42 144, 50, 42

L G H I J K

14 14 14 21, 7 21, 7 21, 28 29 16 33, 27 29 27,

Ecology of B. burgdorferi in East Asia

B. tanukii B. turdi

I. nipponensis A. agrarius A. agrarius

29, 13 29, 13 29, 13

Exact fragment sizes were determined from the sequences. NS, no restriction site. 207

208

K. Miyamoto and T. Masuzawa

afzelii. This conclusion is based on RFLP analysis of the flagellin and 16S rRNA genes. Furthermore, DNA homology studies showed that, while the similarity of the DNA among these isolates ranged from 85 to 99%, DNA similarity with other LB-causing Borrelia spp. was less than 70% (Kawabata et al., 1993). Thus, the unique Borrelia spp. that was isolated from I. ovatus in Japan was subsequently described in 1993 as a new species, Borrelia japonica (Kawabata et al., 1993). While B. burgdorferi s.l. spirochaetes in Japan are most commonly associated with I. persulcatus and I. ovatus, these tick species are by no means the only arthropod hosts. In a study by Nakao and Miyamoto (1993a), several rare Japanese ixodid ticks species such as I. tunuki, I. turdus and I. columnae were found to harbour borrelial spirochaetes. The protein profiles of isolates obtained from I. tanuki and I. turdus were different from those of known species. Of more interest perhaps is that some isolates obtained from I. tanuki and I. turdus have now been classified as new members of the B. burgdorferi s.l. species complex, i.e. Borrelia tanukii and Borrelia turdii, respectively (Fukunaga et al., 1996a). These conclusions were again based on RFLP analysis of 5S–23S rRNA gene intergenic spacer sequence (Masuzawa et al., 1996a,d), ribotyping, PCR targeted to the 16S rRNA gene (Fukunaga et al., 1996b) and DNA–DNA hybridization analysis (Masuzawa et al., 1996c,d). Although strain Am501 from I. columnae (only one isolate thus far obtained) was also initially thought to be unique (Nakao and Miyamoto, 1993a), this strain has subsequently been classified as Borrelia valaisiana (Masuzawa et al., 1996a, 1999; Wang et al., 1997). In summary of this section, the most common B. burgdorferi s.l. species found in Japan are B. garinii and B. afzelii. These two spirochaete species are most commonly isolated from I. persulcatus and LB patients in Japan. Moreover, Japanese B. garinii can be further classified into two subtypes: the Eurasian type (type strain 20047) found in Europe and Asia, and the Asian type (variant strain NT29), found in Asia. Surprisingly, B. burgdorferi s.s. has not yet been found in Japan. The other tick vectors, I. ovatus, I. tanuki and I. turdus, harbour B. japonica, B. tanukii and B. turdii, respectively, while the Japanese strain Am501, which has only been isolated from I. columnae, is now considered to be B. valaisiana.

Reservoir Hosts and Enzootic Transmission Cycles of B. burgdorferi s.l. in Japan Eight species of mammal, two large (C. nippon and Vulpes vulpes ) and six small (Apodemus speciosus ainu, Apodemus argenteus, Clethrionomys rufocanus bedfordia, Clethrionomys rutilus mikado, Eothenmys smithii and Sorex unguiculatus), and four bird species (Turdus chrysolaus, Turdus pallidus, Emberiza spodocehala and Emberiza rustica) have been shown to be infected with B. burgdorferi s.l. in Japan. The spirochaetes, designated as ASF strain (B. garinii), were first isolated in 1989 from the large Japanese field mouse, A. speciosus ainu, captured at Furano (Miyamoto et al., 1991). In further investigations in the Furano and Nemuro areas, spirochaetes were detected from A. speciosus ainu, A. argenteus, C. rufocanus bedfordia, C. rutilus mikado

Ecology of B. burgdorferi in East Asia

209

in Furano, and A. speciosus ainu and C. rutilus mikado in Nemuro. Populations and infection rates were higher in A. speciosus ainu than in any other rodent or vole in either location. Feeding ticks were also more abundant on A. speciosus ainu than A. argenteus. In both locations immature I. persulcatus preferred A. speciosus ainu as a host to any of the other three host species in both areas. For these reasons, it has been suggested that the population of A. speciosus ainu constitutes the most important reservoir of B. burgdorferi s.l. in endemic areas of Hokkaido (Nakao and Miyamoto, 1993b). An additional survey was also carried out involving C. rufocanus bedfordia in Hokkaido, and B. garinii and B. afzelii were isolated from this vole (Ishiguro et al., 1996). B. garinii was taken from A. speciosus in Fukushima, and both B. garinii and B. afzelii were taken from E. smithii in Fukui Prefecture (Ishiguro and Takada, 1996). B. japonica was identified from long-tailed shrews, S. unguiculatus, in Hokkaido (Nakao and Miyamoto, 1993c), from A. speciosus from sites in Aomori, Fukui and Hygo, from A. argenteus in Fukushima, and from E. smithii in Ishikawa and Fukui (Ishiguro et al., 1994; Masuzawa et al., 1995). Large mammals such as red fox and sika deer are the main source of blood for female ticks. The spirochaete B. afzelii was isolated from the skin of a red fox that was infested with all stages of I. persulcatus in Sapporo, Hokkaido (Isogai et al., 1994). A serosurvey of sika deer indicated a high seropositive rate (92.9%) in late October in Hokkaido (Isogai et al., 1996) and B. burgdorferi s.l. was detected in the skin of wild sika deer from east Hokkaido by PCR (Kimura et al., 1995). It is thought that horizontal transmission of LB spirochaetes could occur from infected ticks to agent-free ticks co-feeding on the skin of sika deer (Gern and Rais, 1996). Migratory birds, especially ground-inhabiting birds, also serve as reservoir hosts, and they have been investigated at sites in Nemuro, Hokkaido and Morioka, Iwate Prefecture, in the northern part of Honshu. B. burgdorferi s.l. was detected in larval I. persulcatus feeding on black-faced bunting (E. spodocehala) and red-bellied thrush (T. chrysolaus) (Miyamoto et al., 1993), and B. garinii was directly isolated from three skins and one liver of T. chrysolaus from Nemuro (Miyamoto et al., 1997). Although spirochaetes have been isolated from larvae of I. persulcatus feeding on E. spodocehala, they could not be detected from 60 E. spodocehala from Nemuro (data not shown). Recently, larval I. persulcatus from rustic bunting (E. rustica) and E. spodocehala were examined at Morioka. B. garinii was found from larvae feeding on one E. rustica and on one E. spodocehala (Miyamoto et al., 2000). Attempts to isolate the agent from ten individuals of E. rustica failed (data not shown). Examination of the avian reservoir was also conducted at Fukui Prefecture, in the central part of Honshu. B. garinii was in a larva of I. persulcatus infesting a pale thrush, T. pallidus. In rearing experiments of ticks from the same bird, 38 nymphal H. flava were obtained, all of which moulted to females with only one showing any indication of transstadial transmission (Ishiguro et al., 2000). These results indicate that, as in Europe (Olsen et al., 1995), migratory birds may be involved as reservoirs of B. garinii in Japan. It is considered that these avian reservoirs serve to disperse vector ticks and B. burgdorferi s.l. over long

210

K. Miyamoto and T. Masuzawa

distances worldwide. These birds are infested with larval and/or nymphal ticks from spring to early summer or autumn and are a source of blood for the ticks together with the spirochaetes. The immature ticks become infected with spirochaetes and transstadial transmission to the next stage occurs. Enzootic transmission cycles between birds, rodents and I. persulcatus have been determined by the ribotyping system (Nakao et al., 1994a). Since transovarial transmission of spirochaetes is negligible in I. persulcatus (Nakao and Miyamoto, 1992), enzootic transmission studies have focused on birds and rodents. B. burgdorferi s.l. isolated from larval I. persulcatus fed on migratory birds (Emberiza and Turdus spp.) and rodents (A. speciosus and Clethrionomys spp.) were identified as B. garinii (ribotype II) and B. afzelii (ribotype III), respectively. Ribotype IV, Asiantype B. garinii, can be found in both birds and rodents. Thus, while rodents have a potential to serve as reservoirs for B. afzelii, birds appear to be a specific reservoir host for Eurasian-type B. garinii. This latter fact of is interest when one considers the fact that the Asian-type B. garinii is maintained in both birds and rodents in the wild.

Pathogenicity of Borrelia spp. in Japan Most of the B. burgdorferi s.l. isolates obtained from patients in Japan have been cultured from skin biopsy specimens. Using such specimens, B. garinii ribotype IV (equivalent to Asian-type B. garinii), a few B. garinii ribotype II (equivalent to Eurasian-type B. garinii) and B. afzelii have thus far been identified (Fukunaga et al., 1993c). Experimental transmission studies using tick colonies maintained in the laboratory have revealed that, while I. persulcatus ticks can be readily infected with these Japanese patient-derived B. burgdorferi s.l. isolates after ingestion of the spirochaetes from experimentally infected Mongolian jirds, I. ovatus ticks were not infected (Nakao and Miyamoto, 1994). As mentioned earlier, I. ovatus is one of the most common tick species to infest human habitats in Japan. This tick also specifically transmits B. japonica. None the less, with the possible exception of one suspected case in 1996, clinical cases of bona fide LB due to the transmission of B. japonica through the bite of I. ovatus have not been documented in humans (Masuzawa et al., 1996c). A recent infectivity and arthritis induction study by Kaneda et al. (1998) using B. japonica in severe combined immunodeficiency (scid) mice hints at this spirochaete’s pathogenic potential. In this study, as expected, arthritis induction was observed 12 days after inoculation of scid mice with B. burgdorferi s.s. B. japonica was also able to induce a mild arthritis in scid mice 28 days after inoculation. However, histopathological changes were restricted to the joints. B. japonica also maintained its infectivity and the ability to induce arthritis, albeit with less severe consequences when compared with B. burgdorferi s.s. These results may help to partially explain why the incidence of LB as a result of a bite by I. ovatus is extremely rare in Japan. Since the other rare Japanese tick species, I. tanuki, I. turdus and I. columnae, seldom bite humans, the pathogenic potential of B. tanukii, B. turdi and B. valaisiana in Japan is still unknown.

Ecology of B. burgdorferi in East Asia

211

The findings described here indicate that the principal tick vector of LB in Japan is I. persulcatus. Moreover, the clinical symptoms presented by patients are often limited to erythema migrans with a few people developing facial palsy, arthralgia and lymphocytoma (Nakao et al., 1992b; Hashimoto et al., 1995). We speculate that one of the reasons for this limited clinical spectrum is that Eurasian-type B. garinii, the most dominant Borrelia spp., may have a relatively low virulence in humans. Interestingly, although B. afzelii, a common pathogen in Europe and the species most often implicated in acrodermatitis chronica atrophicans (ACA), is present in Japan, no cases of ACA have been documented. The reason why most LB patients in Japan only present a mild clinical course of the disease is an important unresolved issue.

Distribution of B. burgdorferi s.l. in East Asian Countries While several B. burgdorferi s.l. species have been fairly well characterized in Japan, the species found in other Asian countries have not been extensively studied. To clarify the current distribution of B. burgdorferi s.l. in Asia, we surveyed several East Asian countries. This geographical survey was conducted from 1995 to 2001 and included Far Eastern Russia (Khabarovsk, Vladivostok and Yuzhno-Sakhalinsk), northern China (Yakeshi, Inner Mongolia), southern China (Zhejiang, Sichuan and Anhui provinces), Uighur in western China, Korea, Taiwan and Kinmen island, which is located offshore from Hisamen, Fujian province in China (Fig. 8.2). In these studies, ticks were collected by flagging vegetation in woodland areas and rodents captured with Sherman live traps were collected at the same time. RFLP analysis of the 5S–23S rRNA intergenic spacer sequence was used to determine the Borrelia spp. In addition, the isolates were also characterized by sequence analysis of the intergenic spacer, 16S rRNA and the flagellin gene, and hybridization analysis of genomic DNA. B. burgdorferi s.l. isolated from I. persulcatus collected in the Far Eastern regions of Russia, Inner Mongolia in northeastern China, Uighur in western China and the northern part of South Korea were identified as B. garinii (Eurasian-type and Asian-type) and B. afzelii (Park et al., 1993; Sato et al., 1996; Masuzawa et al., 1997a, 1999; Li et al., 1998; Takada et al., 1998). In contrast, the borrelial isolates obtained from wild rodent species such Apodemus peninsulae captured in Khabarovsk and Inner Mongolia, and A. peninsulae, Apodemus agrarius and Crocidura lasiura from the northern part of South Korea were identified as B. garinii (Asiantype) and B. afzelii. In these areas, B. garinii and B. afzelii appear to be naturally maintained between I. persulcatus and wild mammals. These observations are in good accord with the situation found in the northern part of Japan. In the middle of south Korea (Choongju), where I. persulcatus is not found, Ixodes nipponensis is the main tick vector of these Borrelia spp. Although I. nipponensis, a member of the I. ricinus tick complex, is commonly found in Japan, to

212 K. Miyamoto and T. Masuzawa

Fig. 8.2. Survey sites for vectors and reservoirs of Borrelia burgdorferi sensu lato in East Asian countries from 1995 to 1999 and geographical distribution of B. burgdorferi s.l. in East Asian countries.

Ecology of B. burgdorferi in East Asia

213

our knowledge no isolation of any spirochaete from this tick species has occurred in Japan. However, the genetic differences between isolates from I. persulcatus and I. nipponensis remain to be investigated. Other Korean isolates from I. nipponensis collected in the southern tip of the Korean peninsula have been tentatively classified as belonging to the B. afzelii-related group based on 16S rRNA gene sequences (Kee et al., 1994, 1996). After the initial description of B. valaisiana (Wang et al., 1997), the DNA sequences of several B. valaisiana isolates have been published, so comparison between European and Asian isolates is now possible. We have characterized several strains belonging to B. afzelii-related genomic groups from rodents (Rattus losea, Mus formosanus) captured in Taiwan (Tachung and Kinmen Island) (Masuzawa et al., 2000). B. afzelii-related genomic groups have also been found in Ixodes granulatus and wild rodents (A. agrarius and Niviventer sp.) captured in Zhejiang and Sichuan provinces located in the Yangtze river valley in China (Masuzawa et al., 2001). However, on the basis of the 16S rRNA gene, the 5S–23S rRNA gene intergenic spacer and flagellin gene sequences, Am501, as isolated from the tick I. columnae in Japan, appears to be closely related to B. valaisiana isolated from I. ricinus in Europe (Masuzawa et al., 1996a, 1999). Figure 8.3 shows the phylogenetic tree constructed on the basis of 16S rRNA gene sequences. It can be seen that Korean, Taiwanese and Chinese isolates are clustered with each other, but they are relatively divergent from B. valaisiana in Europe and Am501 found in Japan. Furthermore, OspA gene sequences and protein profiles of major Osps of these spirochaetes were divergent and were not identical to those of the European isolates. These types of borrelia are also present in Thailand, Nepal and the most southern island of Japan, Okinawa (Takada

Fig. 8.3. Phylogenetic tree of Borrelia burgdorferi sensu lato based on 16S rRNA gene sequences. A, isolated from America; C, isolated from China; E, isolated from Europe; J, isolated from Japan; K, isolated from Korea; T, isolated from Taiwan. The bar indicates a sequence divergence of 0.5%.

214

K. Miyamoto and T. Masuzawa

et al., 2001). While it has been recently suggested that some B. burgdorferi s.l. isolates from I. granulatus ticks found in Korea formed a new genomic group (Lee et al., 2000), we feel that these isolates belong to the B. valaisiana-related genomic groups that we have identified. None the less, the classification of B. valaisianarelated species found in Korea, China and Taiwan remains to be determined. On a final note, B. valaisiana found in various European countries is transmitted by I. ricinus ticks (Filipuzzi-Jenny et al., 1993; Livesley et al., 1995; Rijpkema et al., 1995, 1996; Kirstein et al., 1997). Thus, our observations further imply that B. valaisiana has evolved to adapt to a variety of vector ticks in that it has been found in I. granulatus in China, Taiwan and Korea, I. nipponensis in Korea and I. columnae in Japan. However, interesting as these findings are, the pathogenicity of B. valaisiana and related species has yet to be determined. In the survey conducted in the southern part of China (Sichuan and Anhui provinces along the Yangtze river valley), a unique borrelia was isolated from I. ovatus and the wild rodent Niviventer confucianus. These isolates showed reduced similarity of the 16S rDNA sequence to the sequences of previously described Borrelia spp. and could be distinguished by RFLP analysis of the 5S–23S rRNA gene intergenic spacer sequences. The borrelia was morphologically similar to spirochaetes of the B. burgdorferi species complex, but the cells contained fewer periplasmic flagella, a maximum of four, inserted at each end. Based on these findings, the borrelia was named Borrelia sinica, a new member of the of LB-related Borrelia spp. (Masuzawa et al., 2001). This species is distributed in middle China, Thailand and Nepal (unpublished results) and its pathogenicity is still unknown. Vector competence of Haemaphysalis spp. was reported in China (Sheng et al., 1992; Tian et al., 1998). However, no isolate was obtained from Haemaphysalis bispinosa and Haemaphysalis longicorinis collected in China in our survey. Furthermore, in another survey in Nepal and Thailand, cultures made from Haemaphysalis hystricis, Haemaphysalis formonsensis and Haemaphysalis indoflava were all negative. In contrast, a Borrelia sp. was isolated from a few H. flava collected from vegetation by flagging in the Fukui Prefecture, Japan, and was identified as B. japonica (Ishiguro et al., 1994). One nymphal H. flavas fed on T. pallidus, transstadially transmitted B. garinii to the adult stage (Ishiguro et al., 2000). In view of these contradictions, we suggest that Haemaphysalis spp. in Asia are not important as vectors for LB, but may occasionally transmit Borrelia spp. in nature. In summary, the distribution of Borrelia spp. in East Asian countries appears to be wide and varied (Fig. 8.2). B. garinii (Eurasian-type and Asian-type) and B. afzelii transmitted by I. persulcatus ticks are found in Far Eastern Russia, northern and western China, Korea and Japan. On the other hand, the B. valaisianarelated genomic group, which has a more southern distribution in Asia, such as southern China, Taiwan, the southern tip of the Korean peninsula, Thailand and Okinawa in Japan, may be transmitted by I. granulatus. In addition, in Korea this spirochaete may also be transmitted by I. nipponensis. The relationship between Borrelia spp. and vector ticks in East Asian countries is also clarified in Fig. 8.4. I. persulcatus, which is the vector in Russia, northern China, Uighur, the northern part of South Korea and northern Japan, carries

Ecology of B. burgdorferi in East Asia

215

Eurasian-type B. garinii, Asian-type B. garinii and B. afzelii. However, Asian-type B. garinii has not been found in I. ricinus in Europe, and is therefore unique to Asia and I. persulcatus. These Borrelia spp. infect humans and cause LB in their respective countries. However, I. nipponensis, a tick found in mid-Korea (where I. persulcatus is not found), transmits B. garinii and B. afzelii. B. valaisiana is transmitted by I. ricinus, as are B. burgdorferi, B. lusitaniae, Eurasian-type B. garinii and B. afzelii in Europe. In southeast China, Taiwan and Korea, B. valaisiana-related Borrelia spp. are carried by I. granulatus while in Japan this Borrelia sp. is transmitted by I. columnae.

Conclusions We have accumulated much data on the distribution of B. burgdorferi s.l. and their vectors in Japan and parts of eastern Asia. However, there remain many countries and regions in Asia that have yet to be studied. By monitoring the enzootic cycles, transmission and distribution of Borrelia spp. on a worldwide basis, it is hoped that the knowledge gained will lead to the design of better serodiagnosis and prophylaxis regimes. Clinical symptoms presenting in LB patients in Japan are moderate in comparison with those in Europe and the USA, and prophylaxis using vaccines against LB is not justified. At present, the raising of public awareness of this problem and establishment of early diagnosis would seem to be the most effective means of limiting the disease in Japan.

Fig. 8.4. Relationship between Borrelia spp. and vector tick species.

216

K. Miyamoto and T. Masuzawa

Acknowledgements We would like to thank Drs Y. Hashimoto (Asahikawa Medical College, Japan), Y. Sato (previously at Asahikawa Medical College, Japan), F. Sato (Yamashina Institute for Ornithology, Japan), E.I. Korenberg (Gamaleya Institute, Russia), V.I. Volkov (Khabarovsk Antiplague Station, Russia), L.I. Ivanov (Khabarovsk Plague Control Station, Russia), R.N. Liberova (Khabarovsk Regional Center of Sanitary Epidemiological Administration, Russia), N. Takada and Y. Yano (Fukui Medical College, Japan), F. Ishiguro (Fukui Prefectural Institute of Public Health, Japan), H. Fujita (Oohara Institute, Japan), T. Kadosaka (Aichi Medical University, Japan), T. Itoh (Hokkaido Institute of Public Health, Japan), H. Wang and J. Wang (Chinese Medical University, China), K.H. Park (Konkuku University, Korea), M.K. Cho and W.H. Chang (Hallym University, Korea), X.-H. Ma (Zhejiang Institute of Microbiology, China), M.-J. Pan (National Taiwan University, Taiwan) and L.-K. Lin (Tunghai University, Taiwan) for conducting the survey in East Asia countries. We also thank Dr Dennis J. Grab (Johns Hopkins University, School of Medicine, USA, previously at Kurume University School of Medicine, Japan), for critical reading and editing of the final manuscript. This study was supported in part by Grants-in-Aid for International Cooperative Research (Nos 08044310, 08041181 and l0041204) and Grants-in-Aid for Scientific Research (Nos 08670312, 09670294, 11670267) from the Ministry of Education, Science and Culture of Japan.

References Baranton, G., Postic, D., Saint Girons, I., Boerlin, P., Piffaretti, G.-C., Assous, M. and Grimont, P.A.D. (1992) Delineation of Borrelia burgdorferi sensu stricto, Borrelia garinii sp. nov., and group VS461 associated with Lyme borreliosis. International Journal of Systematic Bacteriology 42, 378–383. Bunikis, J., Olsen, B., Fingerle, V., Bonnedahl, J., Wilske, B. and Bergström, S. (1995) Molecular polymorphism of the disease agent Borrelia garinii in Northern Europe is influenced by a novel enzootic Borrelia in the North Atlantic. Journal of Clinical Microbiology 34, 364–368. Canica, M.M., Nato, F., du Merle, L., Mazie, J.C., Baranton, G. and Postic, D. (1993) Monoclonal antibodies for identification of Borrelia afzelii sp.nov. associated with late cutaneous manifes-

tations of Lyme borreliosis. Scandinavian Journal of Infectious Diseases 25, 441–448. Centers for Disease Control (1992) Notifiable diseases report. Morbidity and Mortality Weekly Report 40, 898–899. Filipuzzi-Jenny, E., Blot, M., SchmidBerger, N., Meister-Tuner, J. and Meyer, J. (1993) Genetic diversity among Borrelia burgdorferi isolates: more than three genospecies? Research in Microbiology 144, 295–304. Fukunaga, M. and Koreki, Y. (1996) A phylogenetic analysis of Borrelia burgdorferi sensu lato isolates associated with Lyme disease in Japan by flagellin gene sequence determination. International Journal of Systematic Bacteriology 46, 416–421. Fukunaga, M. and Takahashi, Y. (1994) Pulsed field gel electrophoresis analysis of Borrelia burgdorferi sensu lato isolated

Ecology of B. burgdorferi in East Asia

in Japan and taxonomic implications with Lyme disease spirochetes. Microbiology and Immunology 38, 747–751. Fukunaga, M., Sohnaka, M., Takahashi, Y., Nakao, M. and Miyamoto, K. (1993a) Antigenic and genetic characterization of Borrelia species isolated from Ixodes persulcatus in Hokkaido, Japan. Journal of Clinical Microbiology 31, 1388–1391. Fukunaga, M., Sohnaka, M. and Yanagihara, Y. (1993b) Analysis of Borrelia species associated with Lyme disease by rRNA gene restriction fragment length polymorphism. Journal of General Microbiology 139, 1141–1146. Fukunaga, M., Sohnaka, M., Nakao, M. and Miyamoto, K. (1993c) Evaluation of genetic divergence of borrelial isolates from Lyme disease patients in Hokkaido, Japan, by rRNA gene probes. Journal of Clinical Microbiology 31, 2044–2048. Fukunaga, M., Hamase, A., Okada, K. and Nakao, M. (1996a) Borrelia tanukii sp. nov. and Borrelia turdae sp. nov. found from ixodid ticks in Japan: rapid species identification by 16S rRNA gene-targeted PCR analysis. Microbiology and Immunology 40, 877–881. Fukunaga, M., Hamase, A., Okada, K., Inoue, H., Miyamoto, K. and Nakao, M. (1996b) Characterization of spirochetes isolated from Ixodes tanuki, Ixodes turdus, and Ixodes columnae ticks and sequence comparison with Borrelia burgdorferi sensu lato strains. Applied and Environmental Microbiology 62, 2338–2344. Gern, L. and Rais, O. (1996) Efficient transmission of Borrelia burgdorferi between cofeeding Ixodes ricinus (Acari: Ixodidae). Journal of Medical Entomology 33, 189–192. Hashimoto, Y., Kawagishi, N., Sakai, H., Takahashi, H., Matsuo, S., Nakao, M., Miyamoto, K. and Iizuka, H. (1995) Lyme disease in Japan. Analysis of Borrelia species using rRNA gene restriction fragment length polymorphism. Dermatology 191, 193–198.

217

Hoogstraal, H., Clifford, M., Saito, Y. and Keirans, J.E. (1973) Ixodes (Partipalpiger) ovatus Neuman, subgen. Nov.: identity, hosts, ecology, and distribution (Ixodoidea:Ixodidae). Journal of Medical Entomology 10, 157–164. Ishiguro, F. and Takada, N. (1996) Lyme Borrelia from Ixodes persulcatus and small rodents from northern to central parts of mainland Japan. Medical Entomology and Zoology 47, 183–185. Ishiguro, F., Iida, H., Ma, X.H., Yano, Y., Fujita, H. and Takada, N. (1994) Diversity of Borrelia isolates found in rodent–tick relationship in Fukui Prefecture and some additional areas. Japanese Journal of Sanitation Zoology 45, 141–145. Ishiguro, F., Takada, N., Nakata, K., Yano, Y., Suzuki, H., Masuzawa, T. and Yanagihara, Y. (1996) Reservoir competence of the vole, Clethrionomys rufocanus bedfordiae, for Borrelia garinii or Borrelia afzelii. Microbiology and Immunology 40, 67–69 . Ishiguro, F., Takada, N., Masuzawa, T. and Fukui, T. (2000) Prevalence of Lyme disease Borrelia spp. in ticks from migratory birds on the Japanese mainland. Applied and Environmental Microbiology 66, 982–986. Isogai, E., Isogai, H., Kawabata, H., Masuzawa, T., Yanagihara, Y., Kimura, K., Sakai, T., Azuma, Y., Fujii, N. and Ohno, S. (1994) Lyme disease spirochetes in a wild fox (Vulpes vulpes schrencki) and in ticks. Journal of Wildlife Diseases 30, 439–444. Isogai, E., Isogai, H., Masuzawa, T., Postic, D., Baranton, G., Kamewaka, Y., Kimura, K., Nishikawa, T., Fujii, N., Ishii, N., Ohno, S. and Yamaguti, N. (1996) Borrelia burgdorferi sensu lato in an endemic environment: wild sika deer (Cervus nippon yesoensis) with infected ticks and antibodies. Microbiology and Immunology 40, 13–19. Jaenson, T.G.T. (1991) The epidemiology of Lyme borreliosis. Parasitology Today 7, 39–45.

218

K. Miyamoto and T. Masuzawa

Kaji, K., Miyaki, M., Saitoh, T., Ono, S. and Kaneko, M. (2000) Spatial distribution of an expanding sika deer Cervus nippon population on Hokkaido Island, Japan. Wildlife Society Bulletin 28, 699–707. Kaneda, K., Masuzawa, T., Simon, M.M., Isogai, E., Isogai, H., Yasugami, K., Suzuki, T., Suzuki, Y. and Yanagihara, Y. (1998) Infectivity and arthritis induction of Borrelia japonica on SCID mice and immune competent mice: possible role of galactosylceramide binding activity on initiation of infection. Microbiology and Immunology 42, 171–175. Kawabata, H., Masuzawa, T. and Yanagihara, Y. (1993) Genomic analysis of Borrelia japonica sp. nov. isolated from Ixodes ovatus in Japan. Microbiology and Immunology 37, 843–848. Kawabata, H., Tashibu, H., Yamada, K., Masuzawa, T. and Yanagihara, Y. (1994) Polymerase chain reaction analysis of Borrelia species isolated in Japan. Microbiology and Immunology 38, 591–598. Kawabata, M., Baba, S., Iguchi, K., Yamaguti, N. and Russel, H. (1987) Lyme disease in Japan and its possible incriminated tick vector, Ixodes persulcatus. Journal of Infectious Diseases 156, 854. Kee, S., Hwang, K.J., Oh, H.B. and Park, K.S. (1994) Identification of Borrelia burgdorferi isolated in Korea using outer surface protein A (OspA) serotyping system. Microbiology and Immunology 38, 989–993. Kee, S.H., Yoon, J.H., Oh, H.B., Park, Y.H., Kim, Y.W., Cho, M.K., Park, K.S. and Chang, W.H. (1996) Genetic analysis of Borrelia burgdorferi sensu lato in Korea using genomic hybridization and 16S rRNA gene sequence determination. Microbiology and Immunology 40, 599–605. Kimura, K., Isogai, E., Isogai, H., Kamewaka, Y., Nishikawa, T., Ishii, N. and Fujii, N. (1995) Detection of Lyme disease spirochetes in the skin of naturally infected wild sika deer (Cervus nip-

pon yesoensis). Applied and Environmental Microbiology 61, 1641–1642. Kirstein, F., Rijpkema, S., Molkenboer, M. and Gray, J.S. (1997) The distribution and prevalence of B. burgdorferi genomospecies in Ixodes ricinus ticks in Ireland. European Journal of Epidemiology 13, 67–72. Lee, S.H., Kim, B.J., Kim, J.H., Park, K.H., Yeo, S.J., Kim, S.J. and Kook, Y.H. (2000) Characterization of Borrelia burgdorferi strains isolated from Korea by 16S rDNA sequence analysis and PCR–RFLP analysis of rrf (5S) –rrl (23S) intergenic spacer amplicons. International Journal of Systematic and Evolutionary Microbiology 50, 857–863. Li, M., Masuzawa, T., Takada, N., Ishiguro, F., Fujita, H., Iwaki, A., Wang, H., Wang, J., Kawabata, M. and Yanagihara, Y. (1998) Lyme disease Borrelia species in Northeastern China resemble those isolated from far eastern Russia and Japan. Applied and Environmental Microbiology 64, 2705–2709. Livesley, M.A., Thompson, I.P., Rainey, P.B. and Nuttall, P.A. (1995) Comparison of Borrelia isolated from UK foci of Lyme disease. FEMS Microbiology Letters 130, 151–158. Marconi, R.T. and Garon, C.F. (1992) Development of polymerase chain reaction primer sets for diagnosis of Lyme disease and for species-specific identification of Lyme disease isolates by 16S rRNA signature nucleotide analysis. Journal of Clinical Microbiology 30, 2830–2834. Masuzawa, T., Okada, Y., Beppu, Y., Oku, T., Kawamori, F. and Yanagihara, Y. (1991a) Immunological properties of Borrelia burgdorferi isolated from the Ixodes ovatus in Shizuoka, Japan. Microbiology and Immunology 35, 913–919. Masuzawa, T., Okada, Y., Yanagihara, Y. and Sato, N. (1991b) Antigenic properties of Borrelia burgdorferi isolated from Ixodes ovatus and Ixodes persulcatus in Hokkaido, Japan. Journal of Clinical Microbiology 29, 1568–1573.

Ecology of B. burgdorferi in East Asia

Masuzawa, T., Kawabata, H., Beppu, Y., Miyamoto, K., Nakao, M., Sato, N., Muramatsu, K., Sato, N., Johnson, R.C. and Yanagihara, Y. (1994) Characterization of monoclonal antibodies for identification of Borrelia japonica isolates from Ixodes ovatus. Microbiology and Immunology 38, 393–398. Masuzawa, T., Suzuki, H., Kawabata, H., Ishiguro, F., Takada, N., Yano, Y. and Yanagihara, Y. (1995) Identification of spirochetes isolated from wild rodents in Japan as Borrelia japonica. Journal of Clinical Microbiology 33, 1392–1394. Masuzawa, T., Komikado, T., Iwaki, A., Suzuki, H., Kaneda, K. and Yanagihara. Y. (1996a) Characterization of Borrelia sp. isolated from Ixodes tanuki, I. turdus, and I. columnae in Japan by restriction fragment length polymorphism of rrf (5S) –rrl (23S) intergenic spacer amplicons. FEMS Microbiology Letters 142, 77–83. Masuzawa, T., Wilske, B., Komikado, T., Suzuki, H., Kawabata, H., Sato, N., Muramatsu, K., Sato, N., Isogai, E., Isogai, H., Johnson, R.C. and Yanagihara, Y. (1996b) Comparison of OspA-serotypes for Borrelia burgdorferi sensu lato from Japan, Europe and North America. Microbiology and Immunology 40, 539–545. Masuzawa, T., Yanagihara, Y. and Fujita, H. (1996c) A case of Lyme borreliosis which was suspected to be by Borrelia japonica infection in Shizuoka, Japan. Journal of Japanese Association for Infectious Diseases 70, 264–267 (in Japanese with English summary). Masuzawa, T., Suzuki, H., Kawabata, H., Ishiguro, F., Takada, N. and Yanagihara, Y. (1996d) Characterization of Borrelia spp. isolated from the tick, Ixodes tanuki, and small rodents in Japan. Journal of Wildlife Diseases 32, 565–571. Masuzawa, T., Iwaki, A., Sato, Y., Miyamoto, K., Korenberg, E.I. and Yanagihara, Y. (1997a) Genetic diversity of Borrelia burgdorferi sensu lato isolated

219

in far eastern Russia. Microbiology and Immunology 41, 595–600. Masuzawa, T., Komikado, T., Kaneda, K., Fukui, T., Sawaki, K. and Yanagihara, Y. (1997b) Homogeneity of Borrelia japonica and heterogeneity of Borrelia afzelii and Borrelia tanukii isolated in Japan, determined from OspC gene sequences. FEMS Microbiology Letters 153, 287–293. Masuzawa, T., Kaneda, K., Suzuki, H., Wang, J., Yamada, K., Kawabata, H., Johnson, R.C. and Yanagihara, Y. (1998) Presence of common antigenic epitope in outer surface protein (Osp) A and OspB of Japanese isolates identified as Borrelia garinii. Microbiology and Immunology 40, 455–458. Masuzawa, T., Fukui, T., Miyake, M., Oh, H., Cho, M., Chang, W., Imai, Y. and Yanagihara, Y. (1999) Determination of members of a Borrelia afzelii-related group isolated from Ixodes nipponensis in Korea as Borrelia valaisiana. International Journal of Systematic Bacteriology 49, 1409–1415. Masuzawa, T., Pan, M.-J., Kadosaka, T., Kudeken, M., Takada, N., Yano, Y., Imai, Y and Yanagihara, Y. (2000) Characterization and identification of Borrelia isolates as Borrelia valaisiana in Taiwan and Kinmen Islands. Microbiology and Immunology 44, 1003–1009. Masuzawa, T., Takada, N., Kudeken, M., Fukui, T., Yano, Y., Ishiguro, F., Kawamura, Y., Imai, Y. and Ezaki, T. (2001) Borrelia sinica sp. nov., a Lyme disease-related Borrelia species isolated in China. International Journal of Systematic and Evolutionary Microbiology 51, 1817–1824. Miyamoto, K. and Nakao, M. (1991) Frequent occurrence of human tick bites and monthly fluctuation of ixodid ticks in Hokkaido, Japan. Japanese Journal of Sanitary Zoology 42, 267–269. Miyamoto, K. and Nakao, M. (1994) A study of ixodid tick bite sites on the human body in Hokkaido. Japanese Journal of Sanitary Zoology 45, 79–81.

220

K. Miyamoto and T. Masuzawa

Miyamoto, K., Nakao, M., Sato, N. and Mori, M. (1991) Isolation of Lyme disease spirochetes from an ixodid tick in Hokkaido, Japan. Acta Tropica 49, 65–68. Miyamoto, K., Nakao, M., Uchikawa, K. and Fujita, H. (1992a) Prevalence of Lyme borreliosis spirochetes in ixodid ticks of Japan, with special reference to a new potential vector, Ixodes ovatus (Acari: Ixodidae). Journal of Medical Entomology 29, 216–220. Miyamoto, K., Nakao, M., Fujimoto, K., Yamaguti, N. and Hori, E. (1992b) Detection of Borrelia burgdorferi in ixodid ticks collected from the Chichibu mountainous region of central Honshu, Japan. Japanese Journal of Sanitary Zoology 43, 255–258. Miyamoto, K., Nakao, M., Fujita, H. and Sato, F. (1993) The ixodid ticks on migratory birds in Japan and isolation of Lyme disease spirochetes from birdfeeding ticks. Japanese Journal of Sanitary Zoology 44, 315–326. Miyamoto, K., Sato, Y., Okada, K., Fukunaga, M. and Sato, F. (1997) Competence of a migratory bird, redbellied thrush (Turdus chrysolaus), as an avian reservoir for the Lyme disease spirochetes in Japan. Acta Tropica 65, 43–51. Miyamoto, K., Masuzawa, T. and Kuddeken, M. (2000) Tick collection from wild birds and detection of Lyme disease spirochetes from a new avian reservoir in Japan. Medical Entomology and Zoology 51, 221–226. Nakao, M. and Miyamoto, K. (1992) Negative finding in detection of transovarial transmission of Borrelia burgdorferi in Japanese ixodid ticks, Ixodes persulcatus and Ixodes ovatus. Japanese Journal of Sanitary Zoology 43, 343–345. Nakao, M. and Miyamoto, K. (1993a) Isolation of spirochetes from Japanese ixodid ticks, Ixodes tanuki, Ixodes turdus, and Ixodes columnae. Japanese Journal of Sanitary Zoology 44, 49–52.

Nakao, M. and Miyamoto, K. (1993b) Reservoir competence of the wood mouse, Apodemus speciosus ainu, for the Lyme disease spirochetes, Borrelia burgdorferi, in Hokkaido, Japan. Japanese Journal of Sanitary Zoology 44, 69–84. Nakao, M. and Miyamoto, K. (1993c) Long-tailed shrew, Sorex unguiculatus, as a potential reservoir of the spirochetes transmitted by Ixodes ovatus in Hokkaido, Japan. Japanese Journal of Sanitary Zoology 44, 237–245. Nakao, M. and Miyamoto, K. (1994) Susceptibility of Ixodes persulcatus and I. ovatus (Acari: Ixodidae) to Lyme disease spirochetes isolated from humans in Japan. Journal of Medical Entomology 31, 467–473. Nakao, M., Miyamoto, K., Uchikawa, K. and Fujita, H. (1992a) Characterization of Borrelia burgdorferi isolated from Ixodes persulcatus and Ixodes ovatus ticks in Japan. American Journal of Tropical Medicine Hygiene 47, 505–511. Nakao, M., Miyamoto, K., Kawagishi, N., Hashimoto, Y. and Iizuka, H. (1992b) Comparison of Borrelia burgdorferi isolated from humans and ixodid ticks in Hokkaido, Japan. Microbiology and Immunology 36, 1189–1193. Nakao, M., Miyamoto, K. and Fukunaga, M. (1994a) Lyme disease spirochetes in Japan: enzootic transmission cycles in birds, rodents, and Ixodes persulcatus ticks. Journal of Infectious Diseases 170, 878–882. Nakao, M., Miyamoto, K., Fukunaga, M., Hashimoto, Y. and Takahashi, H. (1994b) Comparative studies on Borrelia afzelii isolated from a patient of Lyme disease, Ixodes persulcatus ticks, and Apodemus speciosus rodents in Japan. Microbiology and Immunology 38, 413–420. O’Connell, S., Granstrom, M., Gray, J.S. and Stanek, G. (1998) Epidemiology of European Lyme borreliosis. Zentralblatt für Bacteriologie 287, 229–240. Olsen, B., Jaenson, T.G.T. and Bergström, S. (1995) Prevalence of Borrelia burgdorferi sensu lato-infected ticks on migrat-

Ecology of B. burgdorferi in East Asia

ing birds. Applied and Environmental Microbiology 61, 3082–3087. Orloski, K.A., Hayes, E.B., Campbell, G.L. and Dennis, D.T. (2000) Surveillance for Lyme disease – United States, 1992–1998. Morbidity and Mortality Weekly Report 49 (SS-3), 1–12. Park, K.H., Chang, W.H. and Schwan, T.G. (1993) Identification and characterization of Lyme disease spirochetes, Borrelia burgdorferi sensu lato, isolated in Korea. Journal of Clinical Microbiology 31, 1831–1837. Postic, D., Assous, M.V., Grimont, P.A.D. and Baranton, G. (1994) Diversity of Borrelia burgdorferi sensu lato evidenced by restriction fragment length polymorphism of rrf (5S) –rrl (23S) intergenic spacer amplicons. International Journal of Systematic Bacteriology 44, 743–752. Postic, D., Korenberg, E., Gorelova, N., Kovalevski, Y.V., Bellenger, E. and Baranton, G. (1997) Borrelia burgdorferi sensu lato in Russia and neighbouring countries: high incidence of mixed isolates. Research in Microbiology 148, 691–702. Rijpkema, S.G.T., Molkenboer, M.J.C.H., Schouls, L.M., Jongejan, F. and Schellekens, J.F.P. (1995) Simultaneous detection and genotyping of three genomic groups of Borrelia burgdorferi sensu lato in Dutch Ixodes ricinus ticks by characterization of the amplified intergenic spacer region between 5S and 23S rRNA genes. Journal of Clinical Microbiology 33, 3091–3095. Rijpkema, S., Golubic, D., Molkenboer, M., Verbeck-de Kruif, N. and Schellekens, J. (1996) Identification of four genomic groups of Borrelia burgdorferi sensu lato in Ixodes ricinus ticks collected in a Lyme borreliosis endemic region of north Croatia. Experimental and Applied Acarology 20, 23–30. Saito, Y. (1959) Studies on ixodid ticks. Part I. On ecology, with reference to distribution and seasonal occurrence of ixodid ticks in Niigata Prefecture,

221

Japan. Acta Medica et Biologica 7, 193–209. Saito, Y. (1960) Studies on ixodid ticks. Part II. On the rearing and life history of three tick species (Haemaphysalis flava, Ixodes japonica and Ixodes persulcatus persulcatus) in Japan (Acarina: Ixodidae). Acta Medica et Biologica 7, 303–321. Sato, Y., Miyamoto, K., Iwaki, A., Masuzawa, T., Yanagihara, Y., Korenberg, E.I., Gorelova, N.B., Ivanov, V.I. and Liberoba, R.N. (1996) Prevalence of the Lyme disease spirochete in Ixodes persulcatus and wild rodents in far east Russia. Applied and Environmental Microbiology 62, 3887–3889. Sheng, Z., Zhang, F. and Wang, J. (1992) Lyme disease spirochetes isolated from Haemaphysalis longicolonis in the forest area of West-North Beijing. Chinese Journal of Vector Biology and Control 3 (suppl. 2), 99–100. Steere, A.C. (1989) Lyme disease. New England Journal of Medicine 321, 586–596. Steere, A.C., Malawista, S.E., Syndman, D.R., Shope, R.E., Andiman, W.A., Ross, M.R. and Steele, F.M. (1977) Lyme arthritis: an epidemic of oligoarticular arthritis in children and adults in three Connecticut communities. Arthritis and Rheumatism 20, 7–17. Takada, N., Ishiguro, F., Fujita, H., Wang, H.-P. and Masuzawa, T. (1998) Lyme disease spirochetes in ticks from northeastern China. Journal of Parasitology 84, 499–504. Takada, N., Fujita, H., Yano, Y., Ishiguro, F., Iwasaki, H. and Masuzawa, T. (2001) First records of tick-borne pathogens, Borrelia, and spotted fever group rickettsiae in Okinawajima island, Japan. Microbiology and Immunology 45, 163–165. Takahashi, Y., Inukai, M., Iguchi, K., Takahashi, I. and Yamamoto, Y. (1997) A case of forest damage caused by sika deer (Cervus nippon) in Iwanazawa experimental plot of the Tokyo university

222

K. Miyamoto and T. Masuzawa

forest in Hokkaido. Transaction Meeting Hokkaido Branch Japan Forestry Society 45, 84–87 (in Japanese). Tian, W., Zhang, Z., Moldenhauer, S., Guo, Y., Yu, Q., Wang, L. and Chen, M. (1998) Detection of Borrelia burgdorferi from ticks (Acari) in Hebei Province, China. Journal of Medical Entomology 35, 295–298. Uchikawa, K., Muramatsu, K., Miyamoto, K. and Nakao, M. (1991) An extensive prevalence of Borrelia burgdorferi, the etiologic agent of Lyme borreliosis, in Nagano Prefecture, Japan. Japanese Journal of Sanitary Zoology 42, 293–299. Wang, G., van Dam, A.P., Le Fleche, A., Postic, D., Peter, O., Baranton, G., de Boer, R., Spajaard, L. and Dankert, J. (1997) Genetic and phylogenetic analysis of Borrelia valaisiana sp. nov. (Borrelia genomic groups VS116 and M19). International Journal of Systematic Bacteriology 47, 926–932. Will, G.S., Jauris-Ileike, S., Schab, E., Busch, U., Rossler, D., Soutschek, E., Wilske, B. and Preac-Mursic, V. (1995)

Sequence analysis of OspA gene shows homogeneity within Borrelia burgdorferi sensu stricto and Borrelia afzelii but reveals major subgroups within the Borrelia garinii species. Medical Microbiology and Immunology 184, 73–80. Wilske, B., Luft, B., Schubach, W.H., Zumstein, G., Jauris, S., Preac-Mursic, V. and Kramer, M.D. (1992) Molecular analysis of the outer surface protein A (OspA) of Borrelia burgdorferi for conserved and variable antibody binding domains. Medical Microbiology and Immunology 181, 191–207. Wilske, B., Preac-Mursic, V., Göbel, U.B., Graf, B., Jauris, S., Schwab, E. and Zumstein, G. (1993) An OspA serotyping system for Borrelia burgdorferi based on reactivity with monoclonal antibodies and OspA sequence analysis. Journal of Clinical Microbiology 31, 340–350. Yamaguti, N. (1989) Human tick bite: variety of tick species and increase of cases. The Saishin-Igaku 44, 903–908 (in Japanese with English summary).

Ecology of Borrelia burgdorferi sensu lato in North America

9

Joseph Piesman Division of Vector-borne Infectious Diseases, National Center for Infectious Diseases, Centers for Disease Control and Prevention, Public Health Service, US Department of Health and Human Services, PO Box 2087, Fort Collins, CO 80522, USA

Introduction The impact of human tick-borne diseases in North America changed dramatically during the last century. Early in the 20th century, tick-borne diseases on the American continent were associated with sparsely populated remote regions of the Rocky Mountains. Rocky Mountain spotted fever, Colorado tick fever and relapsing fever were poorly known diseases associated with frontier regions (Price, 1948). The vast majority of people on the North American continent thought themselves to be free of the risk of tick-borne diseases. These perceptions were altered radically in the late 1960s, or 1969 to be precise. In 1969, Scrimenti (1970) observed an unusual case of erythema chronicum migrans in a Wisconsin resident. During the same year, another unusual case, this time of human babesiosis, was recognized in a resident of Nantucket Island, Massachusetts (Western et al., 1970). Although these two reports were little noticed at the time, they were a portent of greater things to come. In the mid-1970s it became clear that several new tick-borne diseases were becoming increasingly common in highly populated areas of the northeastern USA. In 1977, residents of Lyme, Connecticut (Fig. 9.1), and neighbouring towns were reported to be suffering from an unusual outbreak of arthritis, termed Lyme arthritis (Steere et al., 1977); also in 1977, an ongoing outbreak of human babesiosis was described on Nantucket Island, Massachusetts (Ruebush et al., 1977). Lyme arthritis was soon linked to erythema migrans, a condition long known in Europe, and rapidly becoming associated with the bite of the black-legged tick, Ixodes scapularis. By 1982, a spirochaete had been isolated from these ticks (Burgdorfer et al., 1982) and the aetiologic agent of Lyme disease (Lyme borreliosis) was subsequently named Borrelia burgdorferi (Johnson et al., 1984). This © 2002 CAB International. Lyme Borreliosis: Biology, Epidemiology and Control (eds J. Gray, O. Kahl, R.S. Lane and G. Stanek)

223

224 J. Piesman

Fig. 9.1. Lyme borreliosis risk map for the USA (Fish and Howard, 1999) showing locations (1–16) referred to in the text. (1) Monmouth County, New Jersey; (2) Long Island, New York; (3) Westchester County, New York; (4) Lyme, Connecticut; (5) Nantucket Island, Massachusetts; (6) Monhegan Island, Maine; (7) Long Point, Ontario, Canada; (8) Jasper County, Indiana; (9) British Columbia, Canada; (10) Mendicino County, California; (11) Baja, Mexico; (12) Haulapai Mountain County Park, Mohave County, Arizona; (13) St George, Utah; (14) Larimer County, Colorado; (15) Sapelo Island, Georgia; (16) Camp Lejune, North Carolina.

Ecology of B. burgdorferi in North America

225

nomenclature was modified to B. burgdorferi sensu stricto (s.s.) after the finding that in Europe the spirochaete occured as a complex of genospecies, known as B. burgdorferi sensu lato (s.l.), including B. burgdorferi s.s. (Wilske et al., 1986). Human babesiosis was also linked to I. scapularis transmitting a rodent piroplasm, Babesia microti (Spielman, 1976). With the eye of the biomedical community closely fixed on I. scapularis, additional pathogens were soon found to be associated with this nefarious tick. The agent causing human granulocytic ehrlichiosis (HGE), closely related or identical to Ehrlichia phagocytophila, was found to be transmitted by I. scapularis (Pancholi et al., 1995) and an arbovirus related to an agent that causes human encephalitis, Powassan virus, was also described in this same tick (Telford et al., 1997). It seems prudent to review this sudden temporal and geographical clustering of human disease agents transmitted by ticks in terms of the basic ecology of Lyme borreliosis (LB) spirochaetes and their principal tick vectors in North America.

Tick ecology Systematics I. scapularis is a species that has long been known to science. It was described almost 200 years ago (Say, 1821). Unfortunately, Say’s type specimen cannot be located and he gave as location only ‘United States’ (Nuttall et al., 1911). In an authoritative review of the genus Ixodes in North America, Cooley and Kohls (1945) described the distribution of I. scapularis as being limited primarily to the southern USA, listing 21 records of this tick. This review does, however, cite a report by Cobb (1942) with support from a letter by the noted tick biologist C.N. Smith that placed an isolated population of I. scapularis in a northern location: Cape Cod, Massachusetts. After the summary by Cooley and Kohls (1945), we essentially lost track of I. scapularis for approximately 30 years. Isolated scattered reports placed this tick in various locations in the southern USA from Florida (Rogers, 1953) to Virginia (Kellogg et al., 1971). Northern records included Rhode Island (Hyland and Mathewson, 1961), Maryland and Massachusetts (Kellogg et al., 1971), eastern Long Island, New York (Good, 1973), northern Wisconsin (Jackson and DeFoliart, 1970) and Long Point, Ontario, Canada (Watson and Anderson, 1976) (Fig. 9.1). Few reports on the distribution and biology of I. scapularis prior to 1976 mentioned these ticks as vectors of human or animal disease agents. One exception (Kellogg et al., 1971) listed ‘experimental transmission only’ for bovine anaplasmosis and tularaemia for I. scapularis with no details or references. Then, in 1976, research attention was focused squarely on I. scapularis when nymphs of this tick were identified as the vector of B. microti, the aetiologic agent of human babesiosis on Nantucket Island, Massachusetts (Spielman, 1976). Soon thereafter, nymphal I. scapularis were also implicated as the vector of the yet to be described aetiologic agent of LB (Wallis et al., 1978).

226

J. Piesman

Controversy soon emerged when Spielman et al. (1979) made the bold move of describing the northern Ixodes spp. populations as a new species, Ixodes dammini, separate and distinct from the southern populations of I. scapularis. The grounds for this species designation were primarily based on morphology and geographical distribution. The nymphs, especially, were reported to be different morphologically; a key difference was that I. dammini nymphs contained large triangular posteriorly directed auriculae that arose on the margin of the ventral side of their basis capituli, whereas I. scapularis nymphs contained much smaller auriculae that arose somewhat mesad of the external margin. The result of this characteristic was that I. dammini had pointy auriculae that could be seen with a light microscope and the auriculae of I. scapularis were barely visible. The distribution of I. dammini was reported to include Massachusetts, Connecticut, New Jersey, New York, Rhode Island, Minnesota, Wisconsin and Ontario, Canada; I. scapularis identifications were confirmed from Florida, Georgia, Alabama, Louisiana, Texas, South Carolina, Arkansas, Illinois, Indiana and Iowa (Spielman et al., 1979). An additional distinction was the host preference of immature ticks. Immature I. dammini were reported to feed on a variety of hosts, but to focus on rodents and Peromyscus leucopus (the principal reservoir of B. microti and B. burgdorferi s.s.) in particular. In contrast, I. scapularis immatures were reported to feed predominantly on reptiles, including lizard species that were refractory to infection with these pathogens (Spielman et al., 1984). The literature on the principal vectors of LB spirochaetes in North America generally incorporated the use of I. dammini through the 1980s, until Oliver et al. (1993b) challenged the validity of I. dammini as a species separate and distinct from I. scapularis. Their argument was based on hybridization and assortative mating experiments, morphometric analyses, isoenzymes and host-selection data. The core of their argument for conspecificity was their mating experiments. Reciprocal crosses between I. dammini (Massachusetts) and I. scapularis (Georgia) produced fertile offspring through the F3 generation. No significant decrease in fertility or fecundity occurred in any of the I. scapularis–I. dammini crosses. In contrast, crosses between either I. scapularis or I. dammini with another Ixodes sp. (Ixodes pacificus), produced F1 offspring, but these progeny were sterile. The assortative mating experiments suggested that no mating barriers existed between the Massachusetts (I. dammini) and Georgia (I. scapularis) colonies. Moreover, Oliver et al. (1993b) conducted a morphometric discriminant analysis using 65 measurable characteristics on adults and nymphs of I. scapularis from North Carolina and Georgia and I. dammini from Maryland and Massachusetts, as well as F1 hybrids. I. pacificus was included as an outgroup. Conventional analysis suggested differences in the nymphs of I. scapularis and I. dammini, but these differences were considered to be size-dependent. Gross analysis of the chromosomes and isozymes also showed no significant difference between I. dammini and I. scapularis. A prior experiment (James and Oliver, 1990) was also cited to prove that both I. dammini and I. scapularis fed preferentially on mice as opposed to lizards when presented with a choice. The debate on the species status of I. dammini quickly moved to include the

Ecology of B. burgdorferi in North America

227

new molecular systematic tools. The first in-depth molecular analysis of the validity of I. dammini as a species distinct from I. scapularis focused on ribosomal nucleotide sequences (rDNA), e.g. the internal transcribed spacers ITS1 and ITS2. Wesson et al. (1993) determined the ITS1 and 2 sequences of I. scapularis, I. dammini and I. pacificus. The ITS region for I. pacificus was quite distinct from I. scapularis and I. dammini, but the ITS regions for I. scapularis and I. dammini were similar. Phylogenetic analysis showed that the ITS sequences did not separate I. dammini and I. scapularis. The approach of using ITS sequences for phylogenetic analysis of Ixodes spp. ticks was subsequently challenged by Rich et al. (1997); these authors detected at least 4% nucleotide variation among repeated assays of individual ticks. They concluded that the ITS region should be used with great caution in phylogenetic studies of ticks. In an exhaustive study on the population genetics of I. scapularis (including the putative I. dammini), Norris et al. (1996) concentrated first on the mitochondrial genes 16S and 12S. This analysis clearly distinguished I. scapularis from the two outgroups: I. pacificus and Ixodes ricinus. Among I. scapularis, two distinct clades emerged. The first clade consisted of specimens from Florida to North Carolina and was designated the southern clade. The second clade contained specimens from Mississippi to Massachusetts and was termed the American clade. Mixed populations of the southern and American clades were discovered in Oklahoma, Mississippi, Georgia and North Carolina. The authors concluded that there was 100% support for the monophyletic relationship among all haplotypes of I. scapularis. Interestingly, both clades were present in a single collection from a dog in North Carolina. In order to further test whether these clades were reproductively, if not geographically, isolated, Norris et al. (1996) went on to use random amplified polymorphic DNA PCR (RAPD–PCR) to test for reproductive isolation. Ticks from the two clades consistently grouped together in this RAPD analysis. Overall, Norris et al. (1996) concluded that the two I. scapularis clades were reproductively continuous, and that there was no reproductive isolation among I. scapularis populations. In contrast, a study looking solely at the 16S mitochondrial DNA of I. scapularis and I. dammini observed a similar phenomenon, the existence of two clades, but came to the opposite conclusion (Rich et al., 1995), namely, that a southern clade exists that corresponded to I. scapularis and that a recently expanded clade including northeastern and middle-eastern regions corresponded to I. dammini. A study by the same group looked at the 16S sequences of nine Ixodes spp. from Eurasia and North America (Caporale et al., 1995). They found that all the species were easily separated by > 4% divergence in their 16S nucleotide sequence analysis, except for the two clades of I. scapularis, where divergence was approximately 2%. The pattern of current usage appears to have accepted the conspecificity of I. scapularis and I. dammini, with the vast majority of authors referring to the two clades simply as I. scapularis. An interesting argument for the continued use of the species I. dammini based on epidemiological grounds was presented by Telford (1998). An increased understanding of the biotic and abiotic factors that determine ecological risk of LB spirochaete transmission on a regional level is

228

J. Piesman

certainly needed. The species status of I. scapularis should, however, be determined on the basis of rigorous systematic criteria. A synthesis approach to tick systematics that involves a combination of morphological criteria and multiple genetic loci, as has been recently performed for Rhipicephalus and Boophilus spp. (Beati and Keirans, 2001), may eventually help to resolve some of the conflicts in Ixodes spp. systematics. For the purposes of this chapter, the conspecificity of I. scapularis and I. dammini will be accepted and these ticks will be referred to as I. scapularis.

Habitat In eastern North America, I. scapularis is clearly associated with deciduous forest and habitat containing leaf litter. These ticks are very susceptible to desiccation (Stafford, 1994) and are therefore found in mainland forests where sufficient leaf litter provides a substrate to hold moisture and provide cover from wind, snow and other elements. In Westchester County, New York, the local distribution of I. scapularis was positively correlated with a greenness–wetness index calculated from geographical information system (GIS) data obtained by remote sensing (Dister et al., 1997). A more down-to-earth study conducted in the same area demonstrated that all three active stages of I. scapularis were much more abundant in the woods than other habitats; substantial numbers of ticks were also found in the ecotonal vegetation, while fewer were recovered from ornamental vegetation and even fewer from lawns (Maupin et al., 1991). Those I. scapularis that were found on lawns were in shaded areas that did not dry out during the day. The importance of leaf litter has been demonstrated by experiments where leaf litter was actually removed and tick populations evaluated. This backbreaking experiment showed that I. scapularis populations could be reduced by 72–100% by leaf litter removal (Schulze et al., 1995). In summary, these ticks appear to be woodland species that are closely associated with deciduous forest and leaf litter. The type of forested habitat where I. scapularis is commonly found is quite variable. The classical northeastern hardwood forest in Westchester County, New York, has been described as a mix of oak (Quercus spp.), sugar maple (Acer saccharum), grey birch (Betula populifolia) and tulip poplar (Liriodendron tulipifera) (Daniels et al., 1989). Although I. scapularis clearly abounds in this type of habitat, it is found in many other ecosystems as well. In coastal Massachusetts, the type of vegetation where I. scapularis thrives is more scrub/brush-dominated than mature forest. On Nantucket Island, Massachusetts, habitats with abundant tick populations were characterized by bayberry (Myrica pensylvanica), wild rose (Rosa carolina, Rosa virginiana) and scrub oak (Quercus ilicifolia) (Piesman and Spielman, 1979). In Monmouth County, New Jersey, where mature inland forests are based on sandy soils, a mixture of hardwoods and shrubs are present in areas of high I. scapularis abundance: various oak and maple species, black cherry (Prunus serotina), tulip poplar, American beech (Fagus grandifolia), blackgum (Nyssa sylvati-

Ecology of B. burgdorferi in North America

229

ca) and sweetgum (Liquidambar stryaciflua) represent hardwoods, and the understorey and shrub layer are dominated by saplings of these hardwoods as well as highbush blueberry (Vaccinium corymbosum) and lowbush blueberry (Vaccinium angustifolium), as well as other bushes and greenbriar (Smilax glauca) (Schulze et al., 1998). In New Jersey, an interesting study compared hardwood forests with pine forests containing predominantly pitch pine (Pinus rigida) with mixed hardwood–pine forest, and with old fields where various grasses and shrubs predominated (Schulze et al., 1998). Surprisingly, I. scapularis was highly abundant in hardwood forest, pine forest and mixed hardwood–pine forest habitat. Old field was the only habitat where I. scapularis was consistently rare. In general, one may conclude that I. scapularis is generally associated with hardwood forests containing abundant leaf litter in the eastern USA, but these ticks are quite adaptable and can be found in other habitats such as scrub/brush and pine forests. In Europe, I. ricinus has been reported to thrive in certain grassland areas where sheep graze and trees are totally absent. In North America, I. scapularis has not been described in comparable habitats. An interesting theory has been presented revolving around events called masting, wherein oak trees produce massive amounts of acorn crops every 2–5 years. Researchers in upstate New York conducted an experiment in which a remarkable amount of acorns were placed on small plots (Jones et al., 1998). Tick, mouse and deer populations were monitored in and around these plots. Mouse populations and I. scapularis populations increased on these plots in subsequent years. The authors hypothesized that masting events could influence tick populations from year to year and thereby be used to predict years of high LB risk, which would occur roughly 2 years after a major masting event. Since masting events seem to occur in different areas in the northeastern forest during different years, the scale of these predictions would have to be carefully monitored and the hypothesis tested over many years. Nevertheless, it is an intriguing idea. In the mid-western USA, I. scapularis is generally found in heavily wooded habitats. These areas are often surrounded by vast tracts of land cleared for agriculture in this part of the USA. In general, land dedicated to agricultural crops does not contain I. scapularis (Kitron and Kazmierczak, 1997). An interesting story in Indiana revealed, however, how small a patch of woodlands is needed to allow the establishment of I. scapularis and the natural transmission cycle of B. burgdorferi s.s. Despite concerted efforts to find I. scapularis ticks in Indiana in the 1970s and 1980s, Pinger et al. (1996) could not detect these ticks. A collection of three adult I. scapularis in 1992 and a retrospective report of an LB case in 1985 led these researchers to two sparsely populated counties in northwestern Indiana where the land was largely cleared for agriculture (Jasper and Newton Counties) (Fig. 9.1). Searches led them to a 50 ha woodlot surrounded by cropland. This woodlot had not been cleared because it sat on slightly higher ground than the surrounding flat terrain and within it they found many questing I. scapularis, infested and infected rodents, and B. burgdorferi s.s. infection in adult and nymphal ticks. From this small beginning, I. scapularis has now become well established in the surrounding counties of the northwest corner of Indiana.

230

J. Piesman

The habitat where I. scapularis is found in the southeastern USA is quite diverse. It consists of a mixture of pine forest mixed with hardwoods such as oak and hickory. Dominant trees include long-leaf pine (Pinus palustris), loblolly pine (Pinus taeda), various oaks (Quercus spp.), hickory (Carya spp.) and loblolly bay (Gordonia lasianthus) (Apperson et al., 1993). Included among the understorey are various shrubs including blueberry (Vaccinium crassifolium), inkberry (Ilex glabra), wiregrass (Aristida stricta) and fetterbush (Lyonia lucida). The principal vector of the LB spirochaete in western North America is the western black-legged tick, I. pacificus. There is no systematics controversy surrounding this tick species; an examination of allozymes and mitochondrial genes demonstrated that although there was great haplotype diversity in this species’ range from British Columbia down through Mexico and east to Utah (Fig. 9.1), there is no evidence of genetic isolation among these diverse populations (Kain et al., 1997, 1999). The habitats where I. pacificus thrive are interesting. In north coastal California, the association of I. pacificus has been intensely studied (Li et al., 2000). This area has a Mediterranean climate, with cool wet winters and warm dry summers. Numerous I. pacificus adults have been collected along trails in habitat characterized as Douglas fir forest, north coastal scrub, chaparral and open grasslands in one park, and redwood forest, oak forest, chaparral and open grasslands in another park. Adult I. pacificus were found in these habitats in areas of high brush, with uphill or downhill slope and moderate leaf litter density. Nymphs were found in areas of heavy shade in this study and can generally be found in areas of shade with leaf litter in California. Areas where nymphal I. pacificus were particularly abundant were characterized by the presence of trees such as black oak (Quercus kelloggii), Douglas fir (Pseudotsuga menziesii) and Pacific madrone (Arbutus menziesii) with an understorey of poison oak (Toxicodendron diversiloba) in Cloverdale, California (Clover and Lane, 1995); in Hopland, California, various oaks as well as Pacific madrone, Douglas fir, redwood (Sequoia sempervirens), California bay (Umbellularia californica) and bigleaf maple (Acer macrophyllum) dominated in an area of high nymphal tick abundance (Tälleklint-Eisen and Lane, 1999). Attempts to identify biotic and abiotic factors that dramatically influence the abundance of infected nymphal I. pacificus in Mendicino County, California, were not productive (Tälleklint-Eisen and Lane, 1999). Although I. pacificus is mainly found along the Pacific coast in areas of high rainfall, it can be found in some unusual inland locations where the general environment is quite arid. A population of I. pacificus is extant in the Hualapai Mountain County Park, near Kingman, Mojave County, Arizona (Fig. 9.1). As the name implies, this area borders on one of the driest parts of the world, the Mojave Desert. At sufficient altitude (> 2000 m), however, sky-islands of vegetation containing oak trees and brush, such as the coffee-berry bush (Garrya wrightii) can be found. Snow provides ample moisture and the oak trees provide leaf litter. Numerous I. pacificus were collected here at an altitude of 2345 m (Olson et al., 1992). Similar sky-islands are common in this region and may prove to contain isolated populations of I. pacificus. In another arid region, near St George, Utah, in the southwest corner of Utah, I. pacificus was collected in an

Ecology of B. burgdorferi in North America

231

isolated location called locally ‘Oak Grove’ (Kain et al., 1997, 1999). This population had an extremely restricted diversity of haplotypes when tested with mitochondrial markers, indicating it may be quite isolated and derived from a small number of founder ticks. Finally, immature I. pacificus have been collected from lizards in extremely hot and arid locations in Washington County, in southwestern Utah (J. Piesman, unpublished data). Only the barest of leaf litter could be found along a small stream bed, but nevertheless I. pacificus was found to survive there. In summary, although both I. scapularis and I. pacificus are woodland species associated with high moisture and leaf litter, these ticks can, to a limited degree, extend their distribution to relatively harsh environments.

Hosts The one indispensable piece in the LB puzzle in North America is the presence of deer. In the eastern USA white-tailed deer, Odocoileus virginianus, serve as hosts for all three stages of I. scapularis, so much so, that the tick is generally called the ‘deer tick’, even though the appropriate common name is the black-legged tick. Although all three stages of I. scapularis feed on deer, it is the role of deer as hosts for adult I. scapularis that places it front and centre in the LB story. Soon after I. scapularis was implicated as a vector of human disease, a study was conducted in Massachusetts that sampled deer throughout the year to determine when each tick stage was present on deer. Hundreds of larval, nymphal and adult ticks were found on deer collected from Naushon and Nantucket Islands, Massachusetts (Piesman et al., 1979). Close to 500 adult I. scapularis were removed from a single deer. Numerous publications have since detailed the close association between adult I. scapularis and deer. Other large animals can serve as hosts for adult I. scapularis, e.g. domestic pets (dogs and cats), livestock (cows, horses, sheep, goats, pigs, donkeys) and wild animals such as bears, coyotes, wolves, raccoons and opossums. But infestation levels on these other hosts do not appear to approach those frequently seen on deer. Black bears (Ursus americanus) may be an exception (Kazmierczak et al., 1988); the density of bears, however, is quite low in eastern North America in comparison with deer. Moreover, adult I. scapularis rarely if ever infest smaller mammals such as rabbits, woodchucks, weasels and rodents. Adult I. scapularis also do not infest birds. An interesting question is whether deer are absolutely essential for the establishment and maintenance of I. scapularis. One approach to answering this question is to measure the abundance of I. scapularis in neighbouring locations and correlate the tick abundance with the density of deer. In a macrogeographical study of 13 islands in Essex County, Massachusetts, the density of deer was closely correlated with the abundance of larval I. scapularis on a per island basis (Wilson et al., 1985). Similarly, in a study in Rhode Island, I. scapularis was present on two islands with deer and absent from four islands without deer (Anderson et al., 1987). In a microgeographical study conducted on Long Island, New York, the frequency

232

J. Piesman

with which deer carrying radio-telemetry devices occupied defined quadrants was positively correlated with the number of immature I. scapularis on rodents the following summer (Wilson et al., 1990). In natural areas and parks on Long Island, New York (Duffy et al., 1994), ticks were sampled from 16 locations where deer were abundant and six locations where deer were believed to be absent (a difficult proposition to prove absolutely). The natural areas where deer were abundant had 14-fold more nymphs and 56-fold more larval I. scapularis compared with natural areas where deer were rare or absent. Studies on deer exclusion using fencing have also shown that deer are essential hosts (Stafford, 1993; Daniels and Fish, 1995), as have deer removal experiments (Wilson et al., 1988; Deblinger et al., 1993). There is little doubt that the importance of white-tailed deer as a pivotal host for adult I. scapularis is due, in part, to the superabundance of these animals at the close of the 20th century. In an interesting review of the history of whitetailed deer in North America, McCabe and McCabe (1997) estimate that before European colonization (c. 1500) there were between 24 and 33 million whitetailed deer in what is now the USA. During the period 1850–1900, agricultural expansion and aggressive hunting practices led to the virtual demise of the species. Many counties in New England reported the shooting of the ‘last deer in the county’ from 1895 to 1900. In the latter half of the 20th century, a reversion of agricultural land to forest and lack of hunting pressure have returned the white-tailed deer to some of its former glory. McCabe and McCabe (1997) estimate there are between 16 and 17 million white-tailed deer in the USA at present. What is remarkable is the explosive growth of these populations over a period of 100 years. In Connecticut, for instance, there were probably only about a dozen deer left in 1896. Today there are probably > 75,000 deer in Connecticut (Gregonis, 2000) and as many as 2 million deer in New England and the adjacent mid-Atlantic states. This explosive growth in the white-tailed deer population has no doubt coincided in time with the LB epidemic in northeastern North America. The hypothesis that I. scapularis was once prevalent over much of the eastern deciduous forest, reduced to a few relict populations when deer were virtually eradicated around 1900 and are slowly but surely now returning to their original range as the deer population explodes is intriguing. In western North America, I. pacificus is closely associated with another deer species, namely, the Columbian black-tailed deer, Odocoileus hemionus columbianus (Westrom et al., 1985). This subspecies is among several referred to commonly as ‘mule deer’. In general, mule deer are extremely abundant in areas where I. pacificus are abundant, and adult I. pacificus are found in high numbers on these animals. Although larval and nymphal I. pacificus can be found on deer, other hosts are considered more important for immatures (Westrom et al., 1985). Rodents also play a key role in the ecology of LB in North America, as hosts for immature Ixodes spp. ticks. When attention was first directed at I. scapularis as a vector of human pathogens, Nantucket Island, Massachusetts, was a focal point of early research. This island had a unique paucity of rodent species, with white-footed mice (P. leucopus), meadow voles (M. pennsylvanicus) and short-tailed

Ecology of B. burgdorferi in North America

233

shrews (Blarina brevicauda) as the only rodents present in abundance (Healy et al., 1976). Furthermore, the two main ticks encountered used different hosts for their immature stages, with the American dog tick Dermacentor variabilis mainly associated with meadow voles and I. scapularis mainly associated with white-footed mice. Summarizing numerous studies conducted on Nantucket Island, Spielman et al. (1984) concluded that 91% of larval and 91% of nymphal I. scapularis fed on white-footed mice, with voles and deer feeding far fewer immature I. scapularis. A paradigm emerged from this body of work on Nantucket Island that featured white-tailed deer as the indispensable host for adult I. scapularis and white-footed mice as the indispensable host for immature I. scapularis. The P. leucopus–I. scapularis association proved not to be an absolute requirement. For instance, Monhegan Island, an isolated island off the coast of Maine, was known to have deer and an abundant population of I. scapularis infected with B. burgdorferi s.s., but no white-footed mice (Smith et al., 1993). It transpired that the Norway rat, Rattus norvegicus, had substituted efficiently as the rodent host for immature I. scapularis, thus demonstrating the plasticity of these ticks and the LB disease spirochaete ecological transmission cycle. Deer have now been totally removed from this island in a tick control experiment and it will be interesting to follow the survival of I. scapularis in these unique circumstances. In addition, as mainland locations became the sites of Lyme disease ecological studies, the predominance of P. leucopus as hosts for immature I. scapularis became less clear. In Westchester County, New York, chipmunks, shrews, skunks and raccoons have all been suggested as important alternative mammalian hosts for immature I. scapularis (Fish and Daniels, 1990; Daniels et al., 1991). In the mid-western USA, chipmunks clearly play an important role as hosts for immature I. scapularis. Studies along the Rock River in Illinois revealed that the eastern chipmunk, Tamias striatus, had higher overall levels of infestation with I. scapularis than did P. leucopus and this was especially significant for the nymphal stage of these ticks (Mannelli et al., 1993). The relative importance of these two rodents in the LB spirochaete ecological cycle should be the subject of locally based comparative studies (e.g. Slajchert et al., 1997) examining tick abundance on each host, the infectivity of each host species and their relative abundance. In general, the importance of various rodents as hosts for immature I. scapularis and as reservoirs for B. burgdorferi s.s. requires intense local study. Rodents also play a role as important hosts for immature I. pacificus and in an area of intensive study, the Hopland Field Station in Mendocino County, California, include the deer mouse (Peromyscus maniculatus), the piñon mouse (Peromyscus truei), the brush mouse (Peromyscus boylii), the California kangaroo rat (Dipodomys californicus), the dusky-footed woodrat (Neotoma fuscipes) and the western harvest mouse (Reithrodontomys megalotis). All of these rodents were infested with larval I. pacificus, but nymphal I. pacificus infestation on rodents was rare (Lane and Loye, 1991). The situation with the dusky-footed woodrat (N. fuscipes) is especially interesting. These rodents are often infested with immature Ixodes spp. ticks, but I. pacificus shares this rodent host with another species, Ixodes spinipalpis. For years, I. spinipalpis was identified as Ixodes neotomae until the two species were

234

J. Piesman

synonomized as I. spinipalpis (Norris et al., 1997). A study conducted in southern Oregon confirmed that small numbers of immature I. pacificus and I. spinipalpis infested woodrats as well as deer mice and brush mice (Burkot et al., 1999). In general, rodents are much less significant than lizards as hosts for immature I. pacificus. Large numbers of both larval and nymphal I. pacificus regularly infest the western fence lizard, Sceloporus occidentalis (Lane and Loye, 1989; Tälleklint-Eisen and Eisen, 1999) and the southern alligator lizard, Elgaria multicarinata (Wright et al., 1998). In addition, the range of I. pacificus and S. occidentalis roughly overlaps. Thus, throughout its range, the western fence lizard appears to be the primary host for immature I. pacificus. Clearly, they are more important hosts for immature I. pacificus than are rodents. As will be discussed in the section Borrelia Ecology (below), the fact that western lizards are the primary host for immature I. pacificus and that these reptiles are refractory to infection with B. burgdorferi s.s. has profound implications for the ecological cycle of LB spirochaetes in the western USA. In the northeastern and mid-western USA, where LB is hyperendemic, lizards are generally absent. In fact, rodents are much more abundant than reptiles in the northern USA. To some degree, this helps focus the feeding of immature I. scapularis on rodents that are reservoir-competent for B. burgdorferi s.s. In the southeastern USA, however, lizards are abundant and are the predominant hosts for immature I. scapularis. An extensive comparison of rodents and lizards as hosts for immature I. scapularis was conducted on the coastal plain of North Carolina, on the Marine Corps Base in Camp Lejune (Apperson et al., 1993). Four lizard species were abundantly infested with immature I. scapularis: the southeastern five-lined skink (Eumeces inexpectatus), the ground skink (Scincella lateralis), the broadheaded skink (Eumeces laticeps) and the eastern glass lizard (Ophisaurus ventralis). In contrast, the number of immature I. scapularis infesting rodents in the same location was extremely low. In fact, immature I. scapularis were almost 50-fold more abundant on lizards than rodents in this study. Preserved museum specimens from the southern USA were also found to have large numbers of attached I. scapularis; one Eumeces specimen had > 200 attached I. scapularis (Oliver et al., 1993a). The fact that immature I. scapularis focus on rodents in the northeast and focus on lizards in the southeast has profound implications for the risk of B. burgdorferi s.s. transmission in these two regions (see Borrelia Ecology below). Interest in birds as hosts for immature I. scapularis during the initial phases of LB research focused mainly on the potential role of migrating birds in introducing ticks to new locations (Spielman, 1988; Wilson et al., 1990). Studies in Westchester County, New York, however, revealed that up to 25 resident bird species were frequently infested with immature I. scapularis (Battaly and Fish, 1993). House wrens (Troglodytes aedon), American robins (Turdus migratorius) and the common grackle (Quiscalus quiscula) had notable levels of infestation. Similar results were obtained from various passerine birds in Connecticut (Stafford et al., 1995). In general, ground-inhabiting birds have higher infestation rates than birds that spend most of their time up in the canopy. The fact that immature I. scapularis in different locations may abundantly infest deer, rodents, medium-

Ecology of B. burgdorferi in North America

235

sized mammals, lizards or birds shows the plasticity of the host range of these ticks. Birds have also been studied as hosts for immature I. pacificus. Extensive studies conducted on 24 bird species captured in an oak woodland forest in Yuba County, California, at an elevation of 440–520 m revealed only five immature I. pacificus infesting 138 birds (Manweiler et al., 1990). In contrast, a study conducted in a mixed pine–oak–madrone forest in Placer County, California, revealed that eight ground-dwelling bird species were commonly infested with I. pacificus (Wright et al., 2000). The bird–I. pacificus relationship appears to vary from location to location.

Borrelia Ecology Host complement and reservoir competence Controversy has long surrounded the importance of various animals as reservoirs of B. burgdorferi s.s. in North America. For instance, the role of white-tailed deer as a reservoir serving to infect I. scapularis ticks that feed on them has been hotly debated (Telford et al., 1988; Oliver et al., 1992). Fortunately, an excellent body of work from Europe has helped to shed light on this somewhat confused aspect of LB ecology in North America. Kurtenbach et al. (1998) found that B. burgdorferi s.l. was lysed by certain host sera and that the active component was host complement. By deactivating complement in various ways, it was shown that complement acted directly via the alternative pathway to kill B. burgdorferi s.l., rather than through antibodies. The most borreliacidal serum was that of deer (red deer, Cervus elaphus), which killed B. burgdorferi s.s. as well as all other European genospecies of B. burgdorferi. Curiously, bird complement killed some European genospecies and rodent complement killed other European genospecies, but B. burgdorferi s.s. was relatively resistant to killing by both bird and rodent sera. This basic observation by Kurtenbach et al. (1998) was quickly extended in North America. In one study, B-31, the prototype North American strain of B. burgdorferi s.s. (from Shelter Island, New York), was exposed to sera from whitefooted mice, New Zealand white rabbits and sika deer (Cervus nippon). The exotic sika deer were used as surrogates for white-tailed deer in tick feeding experiments. Whereas mouse and rabbit sera demonstrated negligible borreliacidal activity, exposure to sera from sika deer resulted in almost complete killing of the spirochetes (Nelson et al., 2000). This killing activity was judged to be due to the alternative complement pathway. Another study compared the killing activity of mammalian and reptilian complement. Using a California strain (CA4) of B. burgdorferi s.s., Kuo et al. (2000) demonstrated that complement from the western fence lizard and the southern alligator lizard effectively killed the spirochaetes, whereas sera from deer mice (P. maniculatus) and humans did not. Although definitive testing of sera from white-tailed deer, mule deer and eastern North American lizard species has not been conducted, these observations on

236

J. Piesman

complement are highly suggestive that these deer and lizards serve as zooprophylactic hosts, a concept first proposed by Spielman et al. (1985). This concept stresses the fact that such hosts remove some immature ticks from the proportion of the tick population that will acquire infection with B. burgdorferi s.s.; these hosts thereby serve to lower the overall risk of transmission of LB spirochaetes to people. Of course, the role of deer is a double-edged sword since they are the presumed indispensable host species for the adult stages of I. scapularis and I. pacificus. But lizards appear to be the primary force in driving down tick infection rates with B. burgdorferi s.s. wherever they are found, so much so that in one study the relative abundance of lizards versus mammals was considered as an inverse indicator of infection rate data for assessing risk of transmission of LB spirochaetes (Anon., 1999). If this approach turns out to be justified, it could go a long way towards explaining why risk of LB transmission is so great in the northeastern USA and extremely low in the southern USA. The only observation that contradicts this hypothesis is the finding that certain lizards found in eastern North America under isolated circumstances may actually serve as competent reservoirs for B. burgdorferi s.s. (Levin et al., 1996). Leaving deer and lizards aside as potential reservoir hosts for B. burgdorferi s.s. in North America, we are left with medium-sized mammals, rodents and birds. Although medium-sized mammals can infect I. scapularis with B. burgdorferi s.s. (Fish and Daniels, 1990), the relative scarcity of these hosts compared with rodents and birds in eastern woodlands generally rules them out as important reservoir hosts. The importance of the white-footed mouse (P. leucopus) as a host for immature I. scapularis has been firmly established. Interestingly, nymphal I. scapularis usually feed on these mice in late spring or early summer (May–July), whereas larvae feed on the same mice later in the year (August–September) (Piesman and Spielman, 1979; Carey et al., 1980; Wilson and Spielman, 1985). This feeding pattern is thought to be essential to the efficient transfer of spirochaetes from P. leucopus to I. scapularis since juvenile mice entering the population become infected during the nymphal feeding season and are virtually all infected and maximally infectious at the time larval I. scapularis feed on them. Infected larval I. scapularis are then thought to moult to nymphs and survive the winter (Yuval and Spielman, 1990) and to transmit spirochaetes to a new crop of mice or humans the next year. This interpretation of the maintenance of B. burgdorferi s.s. places a heavy reliance on standard horizontal transmission of spirochaetes from nymphal ticks to rodents and back to larval ticks that then moult to nymphs and renew the cycle. It is supported by experimental observations demonstrating that laboratory-infected P. leucopus are extremely infectious to larval I. scapularis (Donahue et al., 1987) and field-collected P. leucopus from Massachusetts were also highly infectious, infecting 46% of exposed larval I. scapularis in one study (Mather et al., 1989b). In Europe there has been some discussion of co-feeding as an important adjunct to the maintenance of B. burgdorferi s.l. in areas where sheep are the primary host for all three stages of I. ricinus (Ogden et al., 1997). In the northeastern USA, however, the standard horizontal transmission of B. burgdorferi s.s.

Ecology of B. burgdorferi in North America

237

is so efficient that adjuncts like co-feeding do not appear to be necessary to explain the commonly observed 25% infection rate in nymphal I. scapularis and 50% infection rate in adults (Maupin et al., 1991). In the mid-western USA cofeeding may be more important because feeding of larval and nymphal I. scapularis tends to coincide to a greater degree (Jones and Kitron, 2000); however, the situation is not clear-cut in the mid-western USA. In a study area in Illinois, nymphal I. scapularis were mainly focused on chipmunks (T. striatus) as hosts (Mannelli et al., 1993; Slajchert et al., 1997), whereas larvae were mainly focused on white-footed mice (Mannelli et al., 1994). Issues of horizontal versus co-feeding transmission aside, rodents are clearly the primary reservoirs of B. burgdorferi s.s. in eastern North America, with white-footed mice, deer mice, chipmunks, voles, short-tailed shrews and Norway rats all reported to play a role as reservoirs in specific local situations. In California, the dusky-footed woodrat (N. fuscipes) is the principal reservoir of B. burgdorferi s.s. and the California kangaroo rat (D. californicus) is also frequently infected (Brown and Lane, 1992, 1994). Peromyscus spp., however, seem to be less frequent reservoirs of B. burgdorferi s.s. in this region. The genetic diversity of Borrelia spp. found in these rodents makes it difficult to determine the primary reservoir of those spirochaetes causing LB in the western USA (see section on Diversity of Borrelia genospecies in this chapter). Birds clearly have an auxiliary role to play as reservoirs of B. burgdorferi s.s. in North America. Several reports suggest that larval I. scapularis feeding on a variety of naturally infected birds acquired infection with B. burgdorferi s.s. (Stafford et al., 1995); but at least the grey catbird (Dumatella carolinensis) appeared to be incompetent as a reservoir of B. burgdorferi s.s. (Mather et al., 1989a). In contrast, American robins (T. migratorius) infected almost all the I. scapularis experimentally fed on them, although this infectivity waned after 2 months (Richter et al., 2000). A primary importance of birds may prove to be that, when specific control measures are directed towards targeted killing of ticks feeding on reservoir-competent rodents, birds may prove to be a troubling alternative source of spirochaetes to the tick population.

Diversity of Borrelia genospecies B. burgdorferi s.l. has been split into numerous genospecies in Europe and Asia based on minor differences in the 5S–23S and 16S–23S intergenic spacer regions in the ribosomal genes. Spirochaetes in North America were for a long time thought to be more uniform. In fact, B. burgdorferi s.s. was the sole genospecies discussed in the North American literature until Marconi et al. (1995) described isolates from cottontail rabbits and a tick species closely associated with these rabbits (Ixodes dentatus) as a unique genospecies, Borrelia andersonii. It is unknown whether B. andersonii can commonly infect ticks that feed on people, e.g. I. scapularis, and whether human infection with B. andersonii ever occurs. Subsequent to the description of B. andersonii, a diverse collection of spirochaetes

238

J. Piesman

isolated in California were characterized as an additional new species, Borrelia bissettii (Postic et al., 1998). It is now clear that B. bissettii can be found in many regions spanning the USA, including California, Colorado, Illinois and Connecticut. In Larimer County, Colorado, B. bissettii is most commonly found in woodrats (Neotoma mexicana) and the associated tick I. spinipalpis (Schneider et al., 2000). In California, B. bissettii is also closely associated with woodrats and I. spinipalpis, although it occasionally infects the human-biting I. pacificus. Moreover, several strains of Borrelia spp. isolated from Ixodes spp. ticks and rodents in California are clearly separate and distinct from both B. burgdorferi s.s. and B. bissettii. These spirochaetes may be derived from a separate unique cycle involving the rare tick Ixodes jellisoni and its interesting host, the California kangaroo rat (D. californicus) (Lane et al., 1999). Curiously, diversity of B. burgdorferi s.l. found in I. pacificus increases as one heads from north to south, even though the prevalence of infection decreases (T. Schwan, personal communication). Although a large number of isolates from I. scapularis are similarly available along a north–south cline along the Atlantic coast, these isolates are just beginning to be subject to molecular characterization (Lin et al., 2001). Although it is interesting to study and discuss the great diversity of B. burgdorferi s.l. in North America, it is important to remember that the vast majority of isolates from I. scapularis and I. pacificus are indeed B. burgdorferi s.s. Moreover, to my knowledge all wellcharacterized autochthonous human isolates of B. burgdorferi s.l. from North American patients have proved to be B. burgdorferi s.s. (Mathiesen et al., 1997). Current evidence strongly suggests that interrupting the transmission of B. burgdorferi s.s. is where public health effort is needed.

Other Borrelia species in ixodid ticks Argasid ticks in North America are infected with a variety of Borrelia spp., some of which cause relapsing fever. Among the hard ticks, Ixodes spp. are infected with a variety of genospecies of B. burgdorferi s.l. For many years a variety of observers have reported seeing live spirochaetes in an ixodid tick that is extremely common in North America, namely the Lone Star tick, Amblyomma americanum. All three stages of A. americanum aggressively bite people in the southern USA, and an erythema-like rash is commonly associated with the bite of this tick (Campbell et al., 1995). Approximately 1–2% of A. americanum were infected with live spirochaetes across the range of this tick, but these spirochaetes could not be cultured in BSK medium. Finally, a PCR-based molecular study, sequencing the flagellin gene and 16S gene of spirochaetes from an A. americanum from Texas and one from New Jersey, revealed that these ticks were infected with a spirochaete that was intermediate between B. burgdorferi s.l. spirochaetes and relapsing fever spirochaetes (Barbour et al., 1996). The tentative name of Borrelia lonestari was proposed for the A. americanum spirochaete. B. lonestari is most closely related to Borrelia miyamotoi, a spirochaete isolated from Ixodes spp. ticks in Japan, and Borrelia theileri, a spirochaete transmitted by Rhipicephalus and Boophilus

Ecology of B. burgdorferi in North America

239

genera in Africa and the Americas. Rich et al. (2001) have made the interesting suggestion that B. lonestari evolved from B. theileri, which was introduced into the Americas along with Boophilus spp. ticks. Presumably, these spirochaetes then made the jump from Boophilus spp. to Amblyomma spp. when these ticks fed on the same animals, such as deer. It is difficult to test this hypothesis since the spirochaetes involved have not been cultured, the reservoir hosts are unknown and the pathogenicity of the spirochaetes is undetermined. Much work needs to be done in this field. Another closely related spirochaete has been described recently in North America. This spirochaete apparently infects I. scapularis, can be passed from tick to mouse to tick, and is efficiently passed transovarially from female I. scapularis to larval offspring. Although the spirochaete has not been cultured, it has been characterized based on sequence analysis and is most closely related to B. miyamotoi (Scoles et al., 2001). One of the more important consequences of this discovery is the apparent need to re-examine all reports of transovarial transmission of B. burgdorferi s.s. in larval I. scapularis. Several observers have reported collecting questing larvae infected with spirochaetes and have suggested that these spirochaetes were B. burgdorferi s.s. Transovarial transmission of B. burgdorferi s.s. in the laboratory, however, is difficult to reproduce. Molecular characterization of spirochaetes derived from questing larval I. scapularis may soon clarify this situation.

Human Risk Geographical distribution The area of LB transmission risk in North America encompasses areas where I. scapularis and I. pacificus are extant. In the USA, county-by-county surveys established a rough ‘roadmap’ of where these ticks had been found by the early 1990s (Dennis et al., 1998). There is little doubt that I. scapularis is expanding its geographical range. Studies in New York State have shown that these ticks are moving northward up the Hudson River valley. Whereas the tick was only commonly found within New York State on Long Island before the 1980s, it has now established itself virtually throughout New York State. In the mid-western USA, dense populations of I. scapularis were mainly restricted to an area along the border of Minnesota and Wisconsin. Infected ticks are now found in a slowly expanding radius from this central focus. The suggestion that I. scapularis was once found throughout eastern deciduous forests in North America and is now regaining its former range with the re-establishment of deer populations will require close monitoring. Although this review has focused on the ecology of LB in the USA, there are some areas of transmission in Canada. In eastern Canada isolated adult I. scapularis can be found in several places along the Atlantic coast and as far inland as Saskatchewan (Lindsay, R. et al., 1999; Artsob et al., 2000). Self-

240

J. Piesman

sustaining dense populations of I. scapularis infected with B. burgdorferi s.s. appear to be limited, however, to one location along the shore of Lake Erie. This location is Long Point, Ontario, and the I. scapularis population has long been established here (Watson and Anderson, 1976). In a series of excellent ecological studies, the presence of B. burgdorferi-infected ticks in various habitats at Long Point has been documented (Barker et al., 1992; Lindsay, L.R. et al., 1999) and the risk of transmission of B. burgdorferi s.s. to people visiting parkland on Long Point is real. The potential for the spread of I. scapularis to neighbouring locations in Ontario is being closely monitored. Along the Pacific Coast, I. pacificus extends from British Columbia down to Baja, Mexico (Kain et al., 1997, 1999). The relative transmission risk of B. burgdorferi s.s. to people from the northern edge of the distribution of I. pacificus in Canada to the southern edge in Mexico requires further study. In addition, Ixodes uriae infected with Borrelia garinii have been found in association with seabirds in Alaska, but the public health significance of this observation is unknown (Olsen et al., 1995).

Importance of nymphs There seems little doubt that the nymphal stage of I. scapularis is the stage that transmits B. burgdorferi s.s. to the vast majority of LB victims in North America. In the USA, LB is clearly a summer disease with the vast majority of cases reported in July and August (Dennis, 1995). The feeding of nymphal I. scapularis generally starts in May and continues through early July. If one assumes an incubation period for erythema migrans of up to 3 weeks, the time of transmission of B. burgdorferi s.s. in the northeastern and mid-western USA clearly coincides with the activity period of the nymphal tick (Falco et al., 1999). In contrast, adult I. scapularis feed primarily in the colder months, November–April, when few people acquire LB. This may be due in part to reduced outdoor activity during the winter months, but collections of ticks from tick-bite victims in Westchester County, New York, document that people are frequently bitten by both nymphal and adult I. scapularis in regions hyperendemic for LB (Falco et al., 1996). An important reason why nymphs are such good vectors of B. burgdorferi s.s. to tick-bite victims but adult ticks are not lies in the duration of attachment required for successful spirochaete transmission. B. burgdorferi s.s. must first multiply in the midgut of an infected tick and then migrate through the haemolymph to the salivary glands before a sufficient dose of infectious spirochaetes can be transmitted (Ohnishi et al., 2001). This process generally takes about 2 days in both nymphal and adult I. scapularis (Piesman et al., 1991; des Vignes et al., 2001). Since nymphal I. scapularis are quite small, < 2 mm, whereas female I. scapularis are five- to tenfold larger, it is thought that nymphal ticks often escape detection and feed for a sufficient duration of attachment to transmit B. burgdorferi s.s., whereas female ticks are detected and removed quite early in the feeding process. In Westchester County, New York, approximately three-quarters of adult ticks removed from tick-bite victims were removed in the

Ecology of B. burgdorferi in North America

241

first day of feeding, whereas only half of nymphs were detected during the first day (Falco et al., 1996). This difference may be even more pronounced in ticks that are not detected and complete feeding successfully.

Infection in nymphs Infection with B. burgdorferi s.s. among nymphal I. scapularis in the northeastern USA generally varies from 15 to 30% and an estimate of 25% infected is fairly constant when large samples are taken (Piesman et al., 1986; Maupin et al., 1991). Although this spirochaete population shows some genetic variability (Qiu et al., 1997), they are virtually all B. burgdorferi s.s., the genospecies known to be capable of causing human infection in North America. Infection rates in nymphal I. scapularis in the mid-western USA tend to be slightly lower, generally around 15% (Walker et al., 1994). It is clear, however, that in endemic regions in the northeastern and mid-western USA, nymphal I. scapularis are commonly infected with B. burgdorferi s.s., frequently bite people and are extremely competent vectors of spirochaetes to residents and visitors to these areas. In the southern USA the situation is less clear. A large-scale effort to collect ticks removed from people in Georgia and South Carolina turned up approximately 1000 adult ticks, but only a single I. scapularis nymph (Felz et al., 1996). The infection rate of southern populations of nymphal I. scapularis is still unknown because questing nymphal I. scapularis are so hard to collect that sufficient samples for testing are not available. The author spent 1 week dragging for ticks on two islands (Sapelo and St Catherines) off the coast of Georgia where adult I. scapularis are known to be common, at the time when nymphal I. scapularis infesting lizards were at their peak. He managed to collect a dozen questing nymphs during this week and return ten alive to the laboratory; none of them was infected with B. burgdorferi s.l. The author knows of no larger sample of questing nymphal I. scapularis tested from the southern USA. The majority of testing of I. pacificus for B. burgdorferi s.l. has utilized adults. Generally, the infection rate is quite low, 1–2% positive. Surprisingly, when largescale testing of nymphal I. pacificus was conducted in a location in Mendocino County, 14% of nymphal I. pacificus proved to be infected with B. burgdorferi s.l. (Clover and Lane, 1995). Clearly, the majority of LB cases along the Pacific coast are acquired during the spring and summer, when nymphal I. pacificus are active (March–September). But human cases of LB in California do occur throughout the year and some of these may be associated with adult ticks (Lane and Lavoie, 1988; Clover and Lane, 1995; Monsen et al., 1999). A recent study in Mendocino County demonstrated that infection rates in nymphal I. pacificus from two nearby locations varied from 2 to 12% (TälleklintEisen and Lane, 1999). Basic ecological studies are needed at a local and regional level to determine the factors controlling rates of infection with B. burgdorferi s.l. in nymphal tick populations. Such studies are essential to predict risk of LB and to develop a plan to prevent it.

242

J. Piesman

Summary Over the last 25 years our knowledge of the biology and ecology of the ticks that transmit LB spirochaetes in North America has grown tremendously. We have a much better concept of the habitats in which these ticks thrive, the hosts they feed on, the manner in which they acquire and transmit spirochaetes, and the diversity of Borrelia spp. in North America. However, there remains much to be learnt. We do not know the limits of the distribution of vector ticks, whether the removal of essential hosts will eradicate tick populations, and whether B. burgdorferi s.s. is the only Borrelia sp. causing human LB in North America. As we seek the answers to these questions, we must keep in mind that the ultimate goal in accumulating knowledge on the biology and ecology of the ticks that transmit LB spirochaetes is to devise prevention strategies that will stop the transmission of the spirochaetes to people and thus aid the public’s health.

References Anderson, J.F., Johnson, R.C., Magnarelli, L.A., Hyde, F.W. and Myers, J.E. (1987) Prevalence of Borrelia burgdorferi and Babesia microti in mice on islands inhabited by white-tailed deer. Applied and Environmental Microbiology 53, 892–894. Anon. (1999) Recommendations for the use of Lyme disease vaccine. Recommendations of the advisory committee on immunization practice (ACIP). Morbidity and Mortality Weekly Report 48 (RR-7), 1–25. Apperson, C.S., Levine, J.F., Evans, T.L., Braswell, A. and Heller, J. (1993) Relative utilization of reptiles and rodents as hosts by immature Ixodes scapularis (Acari: Ixodidae) in the coastal plain of North Carolina, USA. Experimental and Applied Acarology 17, 719–731. Artsob, H., Maloney, R., Conboy, G. and Horney, B. (2000) Identification of Ixodes scapularis in Newfoundland, Canada. Canadian Communicable Disease Report 26, 133–134. Barbour, A.G., Maupin, G.O., Teltow, G.J., Carter, C.J. and Piesman, J. (1996) Identification of an uncultivable Borrelia species in the hard tick Amblyomma americanum: possible agent of a Lyme disease-

like illness. Journal of Infectious Diseases 173, 403–409. Barker, I.K., Surgeoner, G.A., Artsob, H., McEwen, S.A., Elliott, L.A., Campbell, G.D. and Robinson, J.T. (1992) Distribution of the Lyme disease vector, Ixodes dammini (Acari: Ixodidae) and isolation of Borrelia burgdorferi in Ontario, Canada. Journal of Medical Entomology 29, 1011–1022. Battaly, G.R. and Fish, D. (1993) Relative importance of bird species as hosts for immature Ixodes dammini (Acari: Ixodidae) in a suburban residential landscape of southern New York State. Journal of Medical Entomology 30, 740–747. Beati, L. and Keirans, J.E. (2001) Analysis of the systematic relationships among ticks of the genera Rhipicephalus and Boophilus (Acari: Ixodidae) based on mitochondrial 12S ribosomal DNA gene sequences and morphological characters. Journal of Parasitology 87, 32–48. Brown, R.N. and Lane, R.S. (1992) Lyme disease in California: a novel enzootic transmission cycle of Borrelia burgdorferi. Science 256, 1439–1442.

Ecology of B. burgdorferi in North America

Brown, R.N. and Lane, R.S. (1994) Natural and experimental Borrelia burgdorferi infections in woodrats and deer mice from California. Journal of Wildlife Diseases 30, 389–398. Burgdorfer, W., Barbour, A.G., Hayes, S.F., Benach, J.L., Grunwaldt, E. and Davis, J.P. (1982) Lyme disease – a tickborne spirochetosis? Science 216, 1317–1319. Burkot, T.R., Clover, J.R., Happ, C.M., DeBess, E. and Maupin, G.O. (1999) Isolation of Borrelia burgdorferi from Neotoma fuscipes, Peromyscus maniculatus, Peromyscus boylii, and Ixodes pacificus in Oregon. American Journal of Tropical Medicine and Hygiene 60, 453–457. Campbell, G.L., Paul, W.S., Schriefer, M.E., Craven, R.B., Robbins, K.E. and Dennis, D.T. (1995) Epidemiologic and diagnostic studies of patients with suspected early Lyme disease, Missouri, 1990–1993. Journal of Infectious Diseases 172, 470–480. Caporale, D.A., Rich, S.M., Spielman, A., Telford, S.R. and Kocher, T.D. (1995) Discriminating between Ixodes ticks by means of mitochondrial DNA sequences. Molecular Phylogenetics and Evolution 4, 361–365. Carey, A.B., Krinsky, W.L. and Main, A.J. (1980) Ixodes dammini (Acari: Ixodidae) and associated ixodid ticks in south-central Connecticut, USA. Journal of Medical Entomology 17, 89–99. Clover, J.R. and Lane, R.S. (1995) Evidence implicating nymphal Ixodes pacificus (Acari: Ixodidae) in the epidemiology of Lyme disease in California. American Journal of Tropical Medicine and Hygiene 53, 237–240. Cobb, S. (1942) Tick parasites on Cape Cod. Science 95, 503. Cooley, R.A. and Kohls, G.M. (1945) The genus Ixodes in North America. National Institute of Health Bulletin 184, 1–246. Daniels, T.J. and Fish, D. (1995) Effect of deer exclusion on the abundance of immature Ixodes scapularis (Acari:

243

Ixodidae) parasitizing small and medium-sized mammals. Journal of Medical Entomology 32, 5–11. Daniels, T.J., Fish, D. and Falco, R.C. (1989) Seasonal activity and survival of adult Ixodes dammini (Acari: Ixodidae) in southern New York State. Journal of Medical Entomology 26, 610–614. Daniels, T.J., Fish, D. and Falco, R.C. (1991) Evaluation of host-targeted acaricide for reducing risk of Lyme disease in southern New York State. Journal of Medical Entomology 28, 537–543. Deblinger, R.D., Wilson, M.L., Rimmer, D.W. and Spielman, A. (1993) Reduced abundance of immature Ixodes dammini (Acari: Ixodidae) following incremental removal of deer. Journal of Medical Entomology 30, 144–150. Dennis, D.T. (1995) Lyme disease. Dermatology Clinics 13, 537–551. Dennis, D.T., Nekomoto, T.S., Victor, J.C., Paul, W.S. and Piesman, J. (1998) Reported distribution of Ixodes scapularis and Ixodes pacificus (Acari: Ixodidae) in the United States. Journal of Medical Entomology 35, 629–638. Dister, S.W., Fish, D., Bros, S.M., Frank, D.H. and Wood, B.L. (1997) Landscape characterization of peridomestic risk for Lyme disease using satellite imagery. American Journal of Tropical Medicine and Hygiene 57, 687–692. Donahue, J.G., Piesman, J. and Spielman, A. (1987) Reservoir competence of white-footed mice for Lyme disease spirochetes. American Journal of Tropical Medicine and Hygiene 36, 92–96. Duffy, D.C., Campbell, S.R., Clark, D., DiMotta, C. and Gurney, S. (1994) Ixodes scapularis (Acari: Ixodidae) deer tick mesoscale populations in natural areas: effects of deer, area, and location. Journal of Medical Entomology 31, 152–158. Falco, R.C., Fish, D. and Piesman, J. (1996) Duration of tick bites in a Lyme diseaseendemic area. American Journal of Epidemiology 143, 187–192.

244

J. Piesman

Falco, R.C., McKenna, D.F., Daniels, T.J., Nadelman, R.B., Nowakowski, J., Fish, D. and Wormser, G.P. (1999) Temporal relation between Ixodes scapularis abundance and risk of Lyme disease associated with erythema migrans. American Journal of Epidemiology 149, 771–776. Felz, M., Durden, L.A. and Oliver, J.H. Jr (1996) Ticks parasitizing humans in Georgia and South Carolina. Journal of Parasitology 82, 505–508. Fish, D. and Daniels, T.J. (1990) The role of medium-sized mammals as reservoirs of Borrelia burgdorferi in southern New York. Journal of Wildlife Diseases 26, 339–445. Fish, D. and Howard, C. (1999) Methods used for creating a national Lyme disease risk map. Morbidity and Mortality Weekly Report 48 (RR07), 21–24. Good, N.E. (1973) Ticks of eastern Long island: notes on host relations and seasonal distribution. Annals of the Entomological Society of America 66, 240–243. Gregonis, M. (2000) Aerial deer survey results indicate population is increasing. Connecticut Wildlife 20, 12–13. Healy, G.R., Spielman, A. and Gleason, N. (1976) Human babesiosis: reservoir of infection on Nantucket Island. Science 192, 479–480. Hyland, K.E. and Mathewson, J.A. (1961) The ectoparasites of Rhode Island mammals I. The tick fauna. Journal of Wildlife Diseases 11, 1–14. Jackson, J.O. and DeFoliart, G.R. (1970) Ixodes scapularis Say in northern Wisconsin. Journal of Medical Entomology 7, 124–125. James, A.M. and Oliver, J.H. Jr (1990) Feeding and host preference of immature Ixodes dammini, I. scapularis, and I. pacificus (Acari: Ixodidae). Journal of Medical Entomology 27, 324–330. Johnson, R.C., Schmid, G.P., Hyde, F.W., Steigerwalt, A.G. and Brenner, D.J. (1984) Borrelia burgdorferi sp. nov.: etiologic agent of Lyme disease. International

Journal of Systematic Bacteriology 34, 496–497. Jones, C.G., Ostfeld, R.S., Richard, M.P., Schauber, E.M. and Wolff, J.O. (1998) Chain reactions linking acorns to gypsy moth outbreaks and Lyme disease risk. Science 279, 1023–1026. Jones, C.J. and Kitron, U. (2000) Populations of Ixodes scapularis (Acari: Ixodidae) are modulated by drought at a Lyme disease focus in Illinois. Journal of Medical Entomology 37, 408–415. Kain, D.E., Sperling, F.A.H. and Lane, R.S. (1997) Population genetic structure of Ixodes pacificus (Acari: Ixodidae) using allozymes. Journal of Medical Entomology 34, 441–450. Kain, D.E., Sperling, F.A., Daly, H.V. and Lane R.S. (1999) Mitochondrial DNA sequence variation in Ixodes pacificus (Acari: Ixodidae). Heredity 83, 378–386. Kazmierczak, J.J., Amundson, T.E. and Burgess, E.C. (1988) Borreliosis in freeranging black bears from Wisconsin. Journal of Wildlife Diseases 244, 366–368. Kellogg, F.E., Kistner, T.P., Strickland, R.K. and Gerrish, R.B. (1971) Arthropod parasites collected from white-tailed deer. Journal of Medical Entomology 8, 495–498. Kitron, U. and Kazmierczak, J.J. (1997) Spatial analysis of the distribution of Lyme disease in Wisconsin. American Journal of Epidemiology 145, 558–566. Kuo, M.M., Lane, R.S. and Gicias, P.C. (2000) A comparative study of mammalian and reptilian alternative pathway of complement-mediated killing of the Lyme disease spirochete (Borrelia burgdorferi). Journal of Parasitology 86, 1223–1228. Kurtenbach, K., Sewell, H.S., Ogden, N.H., Randolph, S.E. and Nuttall, P.A. (1998) Serum complement sensitivity as a key factor in Lyme disease ecology. Infection and Immunity 66, 1248–1251. Lane, R.S. and Lavoie, P.E. (1988) Lyme borreliosis in California: acarological, clinical, and epidemiological studies.

Ecology of B. burgdorferi in North America

Annals of the New York Academy of Sciences 539, 192–203. Lane, R.S. and Loye, J.E. (1989) Lyme disease in California: interrelationship of Ixodes pacificus (Acari: Ixodidae), the western fence lizard (Sceloporus occidentalis), and Borrelia burgdorferi. Journal of Medical Entomology 26, 272–278. Lane, R.S. and Loye, J.E. (1991) Lyme disease in California: interrelationship of ixodid ticks (Acari), rodents, and Borrelia burgdorferi. Journal of Medical Entomology 28, 719–725. Lane, R.S., Peavey, C.A., Padgett, K.A. and Hendson, M. (1999) Life history of Ixodes (Ixodes) jellisoni (Acari: Ixodidae) and its vector competence for Borrelia burgdorferi sensu lato. Journal of Medical Entomology 36, 329–340. Levin, M., Levine, J.F., Yang, S., Howard, P. and Apperson, C.S. (1996) Reservoir competence of the southeastern five-line skink (Eumeces inexpectatus) and the green anole (Anolis carolinensis) for Borrelia burgdorferi. American Journal of Tropical Medicine and Hygiene 54, 92–97. Li, X.L., Peavey, C.A. and Lane, R.S. (2000) Density and spatial distribution of Ixodes pacificus (Acari: Ixodidae) in two recreational areas in north coastal California. American Journal of Tropical Medicine and Hygiene 62, 415–422. Lin, T., Oliver, J.H. Jr, Gao, L., Kollars, T.M. Jr and Clark, K.L. (2001) Genetic heterogeneity of Borrelia burgdorferi sensu lato in the southern United States based on restriction fragment length polymorphism and sequence analysis. Journal of Clinical Microbiology 39, 2500–2507. Lindsay, L.R., Mathison, S.W., Barker, I.K., McEwen, S.A. and Surgeoner, G.A. (1999) Abundance of Ixodes scapularis (Acari: Ixodidae) larvae and nymphs in relation to host density and habitat on Long Point, Ontario. Journal of Medical Entomology 36, 243–254. Lindsay, R., Artsob, H., Galloway, T. and Horsman, G. (1999) Vector of Lyme borreliosis, Ixodes scapularis, identified in

245

Saskatchewan. Canadian Communicable Disease Report 25, 81–83. Mannelli, A., Kitron, U., Jones, C.J. and Slajchert, T.L. (1993) Role of the eastern chipmunk as a host for immature Ixodes dammini (Acari: Ixodidae) in northwestern Illinois. Journal of Medical Entomology 30, 87–93. Mannelli, A., Kitron, U., Jones, C.J. and Slajchert, T.L. (1994) Influence of season and habitat on Ixodes scapularis infestation on white-footed mice in northwestern Illinois. Journal of Parasitology 80, 1038–1042. Manweiler, S.A., Lane, R.S., Block, W.M. and Morrison, M.L. (1990) Survey of birds and lizards for ixodid ticks (Acari) and spirochetal infection in northern California. Journal of Medical Entomology 27, 1011–1015. Marconi, R.T., Liveris, D. and Schwartz, I. (1995) Identification of novel insertion elements, restriction fragment length polymorphism patterns, and discontinuous 23S rRNA in Lyme disease spirochetes: phylogenetic analyses of rRNA genes and their intergenic spacers in Borrelia japonica sp. nov. and genomic group 21038 (Borrelia andersonii sp. nov.). Journal of Clinical Microbiology 33, 2427–2434. Mather, T.N., Telford, S.R. III, MacLachlan, A.B. and Spielman, A. (1989a) Incompetence of catbirds as reservoirs for the Lyme disease spirochetes (Borrelia burgdorferi). Journal of Parasitology 75, 66–69. Mather, T.N., Wilson, M.L., Moore, S.I., Ribeiro, J.M.C. and Spielman, A. (1989b) Comparing the relative potential of rodents as reservoirs of the Lyme disease spirochete (Borrelia burgdorferi). American Journal of Epidemiology 130, 143–150. Mathiesen, D.A., Oliver, J.H. Jr, Kolbert, C.P., Tullson, E.D., Johnson, B.J., Campbell, G.L., Mitchell, P.D., Reed, K.D., Telford, S.R. III, Anderson, J.F., Lane, R.S. and Persing, D.H. (1997)

246

J. Piesman

Genetic heterogeneity of Borrelia burgdorferi in the United States. Journal of Infectious Diseases 175, 98–107. Maupin, G.O., Fish, D., Zultowsky, J., Campos, E.G. and Piesman, J. (1991) Landscape ecology of Lyme disease in a residential area of Westchester County, New York. American Journal of Epidemiology 133, 1105–1113. McCabe, T.R. and McCabe, R.E. (1997) Recounting whitetails past. In: McShea, W.J., Underwood, H.B. and Rappole, J.H. (eds) The Science of Overabundance: Deer Ecology and Population Management. Smithsonian Institution Press, Washington, DC, pp. 11–26. Monsen, S.E., Bronson, L.R., Tucker, J.R. and Smith, C.R. (1999) Experimental and field evaluations of two acaricides for control of I. pacificus (Acari: Ixodidae) in northern California. Journal of Medical Entomology 36, 660–665. Nelson, D.R., Rooney, S., Miller, N.J. and Mather, T.R. (2000) Complementmediated killing of Borrelia burgdorferi by nonimmune sera from sika deer. Journal of Parasitology 86, 1232–1238. Norris, D.E., Klompen, J.S.H., Keirans, J.E. and Black, W.C. IV (1996) Population genetics of Ixodes scapularis (Acari: Ixodidae) based on mitochondrial 16S and 12S genes. Journal of Medical Entomology 33, 78–89. Norris, D.E., Klompen, J.S.H., Keirans, J.E., Lane, R.S., Piesman, J. and Black, W.C. IV (1997) Taxonomic status of Ixodes neotomae and I. spinipalpis (Acari: Ixodidae) based on mitochondrial DNA evidence. Journal of Medical Entomology 34, 696–703. Nuttall, G.H.F., Warburton, C., Cooper, W.F. and Robinson, L.E. (1911) Ticks. A Monograph of the Ixodoidea, Part II. Cambridge University Press, London, pp. 133–348. Ogden, N.H., Nuttall, P.A. and Randolph, S.E. (1997) Natural Lyme disease cycles maintained via sheep by co-feeding ticks. Parasitology 115, 591–599.

Ohnishi, J., Piesman, J. and de Silva, A.M. (2001) Antigenic and genetic heterogeneity of Borrelia burgdorferi populations transmitted by ticks. Proceedings of the National Academy of Sciences, USA 98, 670–675. Oliver, J.H. Jr, Stallknecht, D., Chandler, F.W., James, A.M., McGuire, B.S. and Howerth, E. (1992) Detection of Borrelia burgdorferi in laboratory-reared Ixodes dammini (Acari: Ixodidae) fed on experimentally inoculated white-tailed deer. Journal of Medical Entomology 29, 980–984. Oliver, J.H. Jr, Cummins, G.A. and Joiner, M.S. (1993a) Immature Ixodes scapularis (Acari: Ixodidae) parasitizing lizards from the southeastern U.S.A. Journal of Parasitology 79, 684–689. Oliver, J.H. Jr, Owsley, M.R., Hutcheson, H.J., James, A.M., Chen, C., Irby, W.S., Dotson, E.M. and McLain, D.K. (1993b) Conspecificity of the ticks Ixodes scapularis and I. dammini (Acari: Ixodidae). Journal of Medical Entomology 30, 54–63. Olsen, B., Duffy, D.C., Jaenson, T.G., Gylfe, A., Bonnedahl, J. and Bergstrom, S. (1995) Transhemispheric exchange of Lyme disease spirochetes by seabirds. Journal of Clinical Microbiology 33, 3270–3274. Olson, C.A., Cupp, E.W., Luckhart, S., Ribeiro, J.M.C. and Levy, C. (1992) Occurrence of Ixodes pacificus (Parasitoformes: Ixodidae) in Arizona. Journal of Medical Entomology 29, 1060–1062. Pancholi, P., Kolbert, C.P., Mitchell, P.D., Reed, K.D., Dumler, J.S., Bakken, J.S., Telford, S.R. and Persing, D.H. (1995) Ixodes dammini as a potential vector of human granulocytic ehrlichiosis. Journal of Infectious Diseases 172, 1007–1012. Piesman, J. and Spielman, A. (1979) Hostassociations and seasonal abundance of immature Ixodes dammini in southeastern Massachusetts. Annals of the Entomological Society of America 72, 829–832. Piesman, J., Spielman, A., Etkind, P.,

Ecology of B. burgdorferi in North America

Ruebush, T.K. II and Juranek, D.D. (1979) Role of deer in the epizootiology of Babesia microti in Massachusetts, USA. Journal of Medical Entomology 15, 537–540. Piesman, J., Mather, T.N., Donahue, J.G., Levine, J., Campbell, J.D., Karakashian, S.J. and Spielman, A. (1986) Comparative prevalence of Babesia microti and Borrelia burgdorferi in four populations of Ixodes dammini in eastern Massachusetts. Acta Tropica 43, 263–270. Piesman, J., Maupin, G.O., Campos, E.G. and Happ, C.M. (1991) Duration of adult female Ixodes dammini attachment and transmission of Borrelia burgdorferi, with description of a needle aspiration isolation method. Journal of Infectious Diseases 163, 895–897. Pinger, R.R., Timmons, L. and Karris, K. (1996) Spread of Ixodes scapularis (Acari: Ixodidae) in Indiana: collections of adults in 1991–1994 and description of a Borrelia burgdorferi-infected population. Journal of Medical Entomology 33, 852–855. Postic, D., Ras, N.M., Lane, R.S., Hendson, M. and Baranton, G. (1998) Expanded diversity among Californian borrelia isolates and description of Borrelia bissettii sp. nov. (formerly Borrelia group DN127). Journal of Clinical Microbiology 36, 3497–3504. Price, E.G. (1948) Fighting Spotted Fever in the Rockies. Naegele Printing Company, Helena, Montana, 269 pp. Qiu, W.G., Bosler, E.M., Campbell, J.R., Ugine, G.D., Wang, I.N., Luft, B.J. and Dykhuizen, D.E. (1997) A population genetic study of Borrelia burgdorferi sensu stricto from eastern Long Island, New York, suggested frequency-dependent selection, gene flow and host adaptation. Hereditas 127, 203–216. Rich, S.M., Caporale, D.A., Telford, S.R., Kocher, T.D., Hartl, D.L. and Spielman, A. (1995) Distribution of the Ixodes ricinus-like ticks of eastern North

247

America. Proceedings of the National Academy of Sciences, USA 92, 6284–6288. Rich, S.M., Rosenthal, B.M., Telford, S.R., Spielman, A., Hartl, D.A. and Ayala, F.J. (1997) Heterogeneity of the internal transcribed spacer (ITS-2) region within individual deer ticks. Insect Molecular Biology 6, 123–129. Rich, S.M., Armstrong, P.M., Smith, R.D. and Telford, S.R. III (2001) Lone Star tick-infecting borreliae are most closely related to the agent of bovine borreliosis. Journal of Clinical Microbiology 39, 494–497. Richter, D., Spielman, A., Komar, N. and Matuschka, F.R. (2000) Competence of American Robins as reservoir hosts for Lyme disease spirochetes. Emerging Infectious Diseases 6, 133–138. Rogers, A.J. (1953) A study of the ixodid ticks of northern Florida, including the biology and life history of Ixodes scapularis Say (Ixodidae: Acarina). PhD thesis, University of Maryland, USA. Ruebush, T.K., Juranek, D.D., Chisholm, E.S., Snow, P.C., Healy, G.R. and Sulzer, A.J. (1977) Human babesiosis on Nantucket Island. Evidence for self-limited and subclinical infections. New England Journal of Medicine 297, 825–827. Say, T. (1821) An account of the Arachnides of the United States. Journal of the Academy of Natural Science, Philadelphia 2, 59–83. Schneider, B.S., Zeidner, N.S., Burkot, T.R., Maupin, G.O. and Piesman, J. (2000) Borrelia isolates in northern Colorado identified as Borrelia bissettii. Journal of Clinical Microbiology 38, 3103–3105. Schulze, T.L., Jordan, R.A. and Hung, R.W. (1995) Suppresion of subadult Ixodes scapularis (Acari: Ixodidae) following removal of leaf litter. Journal of Medical Entomology 32, 730–733. Schulze, T.L., Jordan, R.A. and Hung, R.W. (1998) Comparison of Ixodes scapularis (Acari: Ixodidae) populations and their habitats in established and emerg-

248

J. Piesman

ing Lyme disease areas in New Jersey. Journal of Medical Entomology 35, 64–70. Scoles, G.A., Papero, M., Beati, L. and Fish, D. (2001) A relapsing fever group spirochete transmitted by Ixodes scapularis ticks. Vector Borne and Zoonotic Diseases 1, 19–31. Scrimenti, R.J. (1970) Erythema chronicum migrans. Archives of Dermatology 102, 104–105. Slajchert, T., Kitron, U.D., Jones, C.J. and Mannelli, A. (1997) Role of the eastern chipmunk (Tamias striatus) in the epizootiology of Lyme borreliosis in northwestern Illinois, USA. Journal of Wildlife Diseases 33, 40–46. Smith, R.P. Jr, Rand, P.W., Lacombe, E.H., Telford, S.R. III, Rich, S.M., Piesman, J. and Spielman, A. (1993) Norway rats as reservoir hosts for Lyme disease spirochetes on Monhegan Island, Maine. Journal of Infectious Diseases 168, 687–691. Spielman, A. (1976) Human babesiosis on Nantucket Island: transmission by nymphal Ixodes ticks. American Journal of Tropical Medicine and Hygiene 25, 784–787. Spielman, A. (1988) Lyme disease and human babesiosis: evidence incriminating vector and reservoir hosts. In: Englund, P.T. and Sher, A. (eds) Biology of Parasitism. A.R. Liss, New York, pp. 147–165. Spielman, A., Clifford, C.M., Piesman, J. and Corwin, M.D. (1979) Human babesiosis on Nantucket Island: description of the vector, Ixodes (Ixodes) dammini, n. sp. (Acarina: Ixodidae). Journal of Medical Entomology 15, 218–234. Spielman, A., Levine, J.F. and Wilson, M.L. (1984) Vectorial capacity of North American Ixodes ticks. Yale Journal of Biology and Medicine 57, 507–513. Spielman, A., Wilson, M.L., Levine, J.F. and Piesman, J. (1985) Ecology of Ixodes dammini-borne human babesiosis and Lyme disease. Annual Review of Entomology 30, 439–460.

Stafford, K.C. III (1993) Reduced abundance of Ixodes scapularis (Acari: Ixodidae) with exclusion of deer by electric fencing. Journal of Medical Entomology 30, 986–996. Stafford, K.C. III (1994) Survival of immature Ixodes scapularis (Acari: Ixodidae) at different relative humidities. Journal of Medical Entomology 31, 310–314. Stafford, K.C. III, Bladen, V.C. and Magnarelli, L.A. (1995) Ticks (Acari: Ixodidae) infesting wild birds (Aves) and white-footed mice in Lyme, CT. Journal of Medical Entomology 32, 453–466. Steere, A.C., Malawista, S.E., Snydman, D.R., Shope, R.E., Andiman, W.A., Ross, M.R. and Steele, F.M. (1977) Lyme arthritis: an epidemic of oligoarticular arthritis in children and adults in three Connecticut communities. Arthritis and Rheumatology 20, 7–17. Tälleklint-Eisen, L. and Eisen, R.J. (1999) Abundance of ticks (Acari: Ixodidae) infesting the western fence lizard, Sceloporus occidentalis, in relation to environmental factors. Environmental and Applied Acarology 23, 731–740. Tälleklint-Eisen, L. and Lane, R.S. (1999) Variation in the density of questing Ixodes pacificus (Acari: Ixodidae) nymphs infected with Borrelia burgdorferi at different spatial scales in California. Journal of Parasitology 85, 824–831. Telford, S.R. III (1998) The name Ixodes dammini epidemiologically justified. Emerging Infectious Diseases 4, 132–134. Telford, S.R. III, Mather, T.N., Moore, S.I., Wilson, M.L. and Spielman, A. (1988) Incompetence of deer as reservoirs of the Lyme disease spirochete. American Journal of Tropical Medicine and Hygiene 39, 105–109. Telford, S.R. III, Armstrong, P.M., Katavolos, P., Foppa, I., Garcia, A.S.O., Wilson, M.L. and Spielman, A. (1997) A new tick-borne encephalitislike virus infecting New England deer ticks, Ixodes dammini. Emerging Infectious Diseases 3, 165–170.

Ecology of B. burgdorferi in North America

des Vignes, F., Piesman, J., Heffernan, R., Schulze, T.L., Stafford, K.C. III and Fish, D. (2001) Effect of tick removal on transmission of Borrelia burgdorferi and Ehrlichia phagocytophila by Ixodes scapularis nymphs. Journal of Infectious Diseases 183, 773–778. Walker, E.D., Smith, T.W., DeWitt, J., Braudo, D.C. and McLean, R.G. (1994) Prevalence of Borrelia burgdorferi in hostseeking ticks (Acari: Ixodidae) from a Lyme disease endemic area in Northern Michigan. Journal of Medical Entomology 31, 524–528. Wallis, R.C., Brown, S.E., Kloter, K.O. and Main, A.J. (1978) Erythema chronicum migrans and Lyme arthritis: field study of ticks. American Journal of Epidemiology 108, 322–327. Watson, T.G. and Anderson, R.C. (1976) Ixodes scapularis Say on white-tailed deer (Odocoileus virginianus) from Long Point, Ontario. Journal of Wildlife Diseases 12, 66–71. Wesson, D.M., McLain, D.K., Oliver, J.H., Piesman, J. and Collins, F.C. (1993). Investigation of the validity of species status of Ixodes dammini (Acari: Ixodidae) using rDNA. Proceedings of the National Academy of Sciences, USA 90, 10221–10225. Western, K.A., Benson, G.D., Gleason, N.N., Healy, G.R. and Schultz, M.G. (1970) Babesiosis in a Nantucket resident. New England Journal of Medicine 283, 854–856. Westrom, D.R., Lane, R.S. and Anderson, J.R. (1985) Ixodes pacificus (Acari: Ixodidae): population dynamics and distribution on Columbian black-tailed deer (Odocoileus hemionus columbianus). Journal of Medical Entomology 22, 507–511. Wilske, B., Preac-Mursic, V., Schierz, G. and von Busch, K. (1986) Immunochemical and immunological analysis of

249

European Borrelia burgdorferi strains. Zentralblatt für Bakteriologie und Hygiene A 263, 92–102. Wilson, M.L. and Spielman, A. (1985) Seasonal activity of immature Ixodes dammini (Acari: Ixodidae). Journal of Medical Entomology 22, 408–414. Wilson, M.L., Adler, G.H. and Spielman, A. (1985) Correlation between deer abundance and that of the deer tick Ixodes dammini (Acari: Ixodidae). Annals of the Entomological Society of America 78, 172–176. Wilson, M.L., Telford, S.R. III, Piesman, J. and Spielman, J. (1988) Reduced abundance of immature Ixodes dammini (Acari: Ixodidae) following removal of deer. Journal of Medical Entomology 25, 224–228. Wilson, M.L., Ducey, A.M., Litwin, T.S., Gavin, T.A. and Spielman, A. (1990) Microgeographic distribution of immature Ixodes dammini ticks correlated with that of deer. Medical and Veterinary Entomology 4, 151–159. Wright, S.A., Lane, R.S. and Clover, R.S. (1998) Infestation of the southern alligator lizard (Squamata: Anguidae) by Ixodes pacificus (Acari: Ixodidae) and its susceptibility to Borrelia burgdorferi. Journal of Medical Entomology 35, 1044–1049. Wright, S.A., Thompson, M.A., Miller, M.J., Knerl, K.M., Elms, S.L., Karpowicz, J.C., Young, J.F. and Kramer, V.L. (2000) Ecology of Borrelia burgdorferi in ticks (Acari: Ixodidae), rodents, and birds in the Sierra Nevada foothills, Placer County, California. Journal of Medical Entomology 37, 909–918. Yuval, B. and Spielman, A. (1990) Duration and regulation of the development cycle of Ixodes dammini (Acari: Ixodidae). Journal of Medical Entomology 27, 196–201.

Epidemiology of Lyme Borreliosis

10

David T. Dennis and Edward B. Hayes Division of Vector-borne Infectious Diseases, National Center for Infectious Diseases, Centers for Disease Control and Prevention, Public Health Service, US Department of Health and Human Services, PO Box 2087, Fort Collins, CO 80522, USA

Introduction Lyme borreliosis (LB) is the most commonly reported arthropod-borne disease in North America and Europe, accounting for tens of thousands of new cases yearly in both regions (O’Connell et al., 1998; Orloski et al., 2000). The disease has long been endemic in Europe, where its principal clinical syndromes were described more than 50 years ago, and where early indirect evidence suggested a tick-transmitted infection responsive to antibiotics (Weber, 2001). In the USA, cases of erythema migrans (EM) were initially recognized in the 1960s and 1970s, and epidemiological evidence suggests that LB became endemic at several foci at that time (Steere, 1989; Spielman, 1994). Molecular examinations of archived materials show that natural cycles of Borrelia burgdorferi sensu lato (s.l.), the agent of LB, were present in northeastern coastal areas in the late 19th century, and they have probably been present in North America for millennia (Persing et al., 1990; Marshall et al., 1994). The seminal investigations of an outbreak of arthritis in coastal Connecticut in the mid-1970s quickly established the principal epidemiological and ecological features of LB: a multi-system inflammatory disease affecting children and adults of both sexes; clustering of cases within townships and even by road and household; onsets of illness mostly occurring in summer months; association of cases with residence in wooded areas where tick-infested mice and deer were abundant; and linkage of illness to antecedent tick bites (Steere et al., 1977, 1978). In the USA, the annual number of reported cases of LB has increased more than 30-fold since B. burgdorferi was first identified in 1982 (Centers for Disease Control and Prevention, 2001). The increase has been associated with an increasing density and distribution of vector ticks (Godsey et al., 1987; White et al., 1991; © 2002 CAB International. Lyme Borreliosis: Biology, Epidemiology and Control (eds J. Gray, O. Kahl, R.S. Lane and G. Stanek)

251

252

D.T. Dennis and E.B. Hayes

Stafford et al., 1998) and changing demographics, which place more persons at risk of infective tick bites (Bowen et al., 1984). The consequence has been a rise in case rates in established endemic foci (Petersen et al., 1989; White et al., 1991) and an expanding endemic range that now includes parts or all of more than 15 states (Dennis et al., 1998; Orloski et al., 2000). The early rapid rise in cases and the geographical expansion have justified the designation of LB as one of the most rapidly emerging infections in the USA (Institute of Medicine, 1992). More recently, spread of the disease has been characterized by a slower but steady increase of incidence among counties on the inland border of LB-endemic areas (Centers for Disease Control and Prevention, unpublished). In Eurasia, disease incidence and geographical range have been relatively stable, and increases in numbers of reported cases are thought to be due mainly to a greater awareness and enhanced surveillance (O’Connell et al., 1998; Strle, 1999). An understanding of the epidemiology of LB requires a thorough knowledge of the complex natural history of B. burgdorferi s.l., its tick vectors and vertebrate hosts, as well as those human activities and behaviours that place persons at risk of infection. LB is, epidemiologically, firstly a disease of place, and maps of the distribution of cases clearly identify endemic areas and their relationship to the distribution of known vector ticks and other biological variables, such as availability of suitable vertebrate reservoirs of infection and maintenance hosts for ticks (Spielman et al., 1985; Fish, 1995; Wilson, 1998; Centers for Disease Control and Prevention, 1999). Descriptive studies have yielded consistent demographic profiles of cases and trends in disease distribution (Petersen et al., 1989; White et al., 1991; Alpert et al., 1992; Berglund et al., 1995; de Mik et al., 1997; Huppertz et al., 1999; Strle, 1999; Orloski et al., 2000). Analytical epidemiological studies show associations between incident disease and various environmental and behavioural factors that result in increased risk of exposure to infective ticks (Ley et al., 1995; de Mik et al., 1997; Cromley et al., 1998; Orloski et al., 1998; Belongia et al., 1999). Epidemiological studies have, however, poorly characterized various factors that place some persons with endemic exposures at greater risk than others, and have not clearly shown beneficial effects of various measures of personal protection. This chapter provides a synopsis of biological factors relevant to the epidemiology of LB (for details, see Chapters 6, 7, 8 and 9); it outlines the known distribution and determinants of the disease, using information provided by surveillance data, surveys and epidemiological studies, and it briefly addresses some cost- and prevention-effectiveness evaluations.

The Agent and its Vector Associations LB is caused by infection with several spirochaete genospecies of the grouping B. burgdorferi s.l. – a diversity that has important epidemiological and ecological implications (Gern and Humair, 1998; Wang et al., 1999). Transmission occurs as a result of bites by infective hard ticks of the Ixodes ricinus species complex. In

Epidemiology of Lyme Borreliosis

253

North America, B. burgdorferi sensu stricto (s.s.) is the only known cause of human LB and, as far as is known, it is transmitted to humans solely by Ixodes scapularis (the black-legged tick) in eastern regions and Ixodes pacificus (the western blacklegged tick) in the far western region (Spielman et al., 1985; Lane et al., 1991). These ticks have a discrete, non-overlapping distribution (Dennis et al., 1998). In Europe, there is a wider diversity of human-infecting B. burgdorferi s.l. The several genospecies known to cause disease are, in order of decreasing disease-causing frequency, Borrelia afzelii, Borrelia garinii and B. burgdorferi s.s. (Saint Girons et al., 1998). Eurasian strains are transmitted by I. ricinus (the European sheep tick) and Ixodes persulcatus (the taiga tick), which have different but overlapping ranges. In eastern Russia and in Asia, only B. afzelii and B. garinii are recognized human pathogens, transmitted in these regions by I. persulcatus ticks (Korenberg, 1994; Postic et al., 1997; Li et al., 1998). Other related Borrelia spp. are less easily classified. In the USA, Borrelia lonestari sp. nov., a hitherto uncultivable borrelia, was identified by DNA analysis in the midgut contents of Amblyomma americanum (lone star) ticks (Barbour et al., 1996), and it has been shown that bites by these ticks are associated in some instances with an illness characterized by mild constitutional symptoms and a rash similar to EM (Campbell et al., 1995; Barbour et al., 1996; Kirkland et al., 1997; Felz et al., 1999; James et al., 2001). An unnamed agent newly identified in I. scapularis ticks in the eastern USA (Scoles et al., 2001) is, like B. lonestari, more closely related to Borrelia theileri and the relapsing fever borreliae than to B. burgdorferi s.l. The pathogenic status of Borrelia bissettii, an agent associated with Ixodes spinipalpis ticks and woodrats in the western USA (Schneider et al., 2000), is ambiguous. Although genetically close to B. burgdorferi s.s., B. bissettii has not been related epidemiologically or clinically with human infection in the USA, but has been found in I. ricinus ticks in Europe and linked to human LB in Slovenia (Picken et al., 1996). Borrelia andersoni, a spirochaete associated with Ixodes dentatus ticks and rabbits in the southern and eastern USA, is not known to cause human borreliosis. In Eurasia, there are several Borrelia spp. vectored by Ixodes spp. ticks that are closely related to B. burgdorferi s.s. but have separate natural histories and are not known to infect humans; examples include Borrelia valaisiana and Borrelia lusitaniae in Europe, and Borrelia japonica and Borrelia miyamotoi in Japan (see Gern and Humair, Chapter 6 and Miyamoto and Masuzawa, Chapter 8).

Reservoirs of Infection (see Kurtenbach et al., Chapter 5) In North America, various rodents serve as the principal reservoirs of infection of B. burgdorferi s.s, predominantly Peromyscus leucopus, the white-footed mouse, in the eastern LB-endemic regions, and wood rats (Neotoma spp.) and kangaroo rats in the far-western endemic regions (Spielman et al., 1985; Brown and Lane, 1996). Several bird species are competent reservoirs in eastern regions, such as robins (Turdus migratorius) (Richter et al., 2000). In Europe, a recent review identified nine small mammals (including several mice, the bank vole and shrews), seven

254

D.T. Dennis and E.B. Hayes

medium-sized mammals (especially squirrels) and a number of birds as competent hosts of human pathogenic strains (Gorelova et al., 1995; Gern et al., 1998). The Norway rat, Rattus norvegicus, has been shown to be a competent reservoir host in both North America and Europe (Smith et al., 1993; Matuschka et al., 1997). Various feral rodents are known reservoirs of infection in eastern regions of the former USSR and in China, and rodents and birds have been shown to be important reservoir hosts in Japan (Nakao et al., 1996).

Transmission of Infection to Humans Mode of transmission Vector ticks transmit infection by cutaneous inoculation of spirochaete-infected saliva. Experimental and epidemiological studies indicate that this occurs only after the tick begins taking a blood meal, which for I. scapularis and I. pacificus usually begins after 36–48 h of tick attachment (Sood et al., 1997; des Vignes et al., 2001). There is evidence that transmission may sometimes occur earlier after I. ricinus or I. persulcatus attachment (Kahl et al., 1998). In both North America and Europe, the nymphal stage of vector ticks is thought to be responsible for the majority of LB cases (Falco et al., 1999; Robertson et al., 2000a). Nymphal ticks often have high infection rates (15–40% in highly endemic areas); they quest in the spring and summer months when people are most likely to engage in outdoor activities and they are so small that they are difficult to detect on clothing or body surfaces. Questing larvae, on the other hand, are rarely infected, principally seek small vertebrate hosts in leaf litter and are unimportant in transmitting infection to humans. Although adult ticks may have a high prevalence of infection, in the USA they usually quest in the period of late autumn to early spring, when people are less apt to be exposed to tick habitat and more likely to wear protective clothing. Furthermore, adult ticks are large enough to be often noticed and removed before feeding. However, in the former USSR and Asia, where the main vector is I. persulcatus, most cases are probably caused by adults because nymphs of this species rarely bite people (Korenberg, 1994). Although B. burgdorferi s.l. has been isolated from the mouthparts of various biting insects (Magnarelli and Anderson, 1988), there is no epidemiological evidence that arthropods other than Ixodes spp. ticks are important in transmission of infection to humans. LB is not contagious; there is no good evidence that infection spreads between humans by sexual contact, through breast milk or through blood transfusions. Transplacental infection in humans occurs readily in other spirochaetal diseases, and has been suggested for LB. Rare silver-stained spirochaetal structures have been described in fetal and neonatal tissues in association with adverse outcomes of pregnancy in residents of LB-endemic areas (Schlesinger et al., 1985); however, histological materials stained with silver are notoriously difficult to interpret, and transplacental transmission of B. burgdorferi s.l. in humans has not been

Epidemiology of Lyme Borreliosis

255

supported by cultural isolation, by immunohistochemical staining or by DNA characterization of spirochaetes obtained from fetal tissues. In the one report of cultural recovery of B. burgdorferi s.s. from fetal tissue, the culture was contaminated by a Bacillus sp. and the identification of B. burgdorferi s.l. was made indirectly by immunofluorescence staining (MacDonald et al., 1987). If maternal infections do result in transplacental infection and fetal or neonatal death, these occurrences must be rare.

Dynamics of infection in humans and classification of disease LB of humans is a multi-stage, multi-system, inflammatory disease, and patients can usually be categorized as having early localized, early disseminated or late persistent infection (Steere, 1989) (see Stanek et al., Chapter 1). The incubation period of LB is days to weeks from infective exposure to earliest manifestations, such as EM rash, lymphadenitis and mild constitutional symptoms. More than 80% of patients develop a recognized EM rash that appears 3–30 days after a tick bite (median 7–10 days) (Nadelman and Wormser, 1995). Not all patients, however, recognize EM or other manifestations of early acute infection. In a large prospective study of the efficacy of vaccine against LB, persons receiving placebo experienced approximately one incident of asymptomatic infection for every four definite cases (Steere et al., 1998). In patients who do not recognize early disease, a latency period of weeks or months may elapse before the diagnosis is triggered by the appearance of later-stage manifestations such as arthritis, various neurological syndromes or acrodermatitis chronica atrophicans (ACA). The period from exposure to the development of laboratory evidence of infection may be several days or weeks, depending on the test employed. B. burgdorferi s.l. can be isolated in culture from early EM lesions (up to 80% success from skin biopsies in untreated patients), or from the plasma of untreated EM patients (up to 50% success) (Wormser et al., 2000). The usual duration of infection in untreated patients is unknown; occasional isolations have been made of B. burgdorferi s.l. from skin, cerebrospinal fluid, myocardium and synovial fluid, months and even years after onset of symptoms (Snydman et al., 1986; Stanek et al., 1990; Pfister et al., 1991; Strle et al., 1995). IgM antibodies directed against B. burgdorferi s.l. antigens usually appear within 3 weeks of infective exposure, and peak between 4 and 6 weeks post-exposure, when IgG antibodies are also likely to be present. Although antibodies tend to wane, they may remain detectable for months or years, with or without treatment, and serology has no role in measuring the response to treatment (Nadelman and Wormser, 1998). Considerable controversy exists over whether chronic active infection can persist in the absence of detectable antibodies directed against B. burgdorferi s.l. Seronegative, disseminated infection has been reported, in which T lymphocyte reactivity was the only immunological indicator of infection (Dattwyler et al., 1988); this phenomenon is, however, probably quite rare. Similarly controversial

256

D.T. Dennis and E.B. Hayes

is the question of persistent infection following one or more appropriate courses of antimicrobial drugs. In the USA, reports describe patients with objective evidence of chronic neuroborreliosis who require more than one course of treatment to bring about a satisfactory resolution of illness, but these patients appear to be the exception rather than the rule (Logigian et al., 1990). Persons with chronic and refractory arthritic manifestations of LB are more likely than controls to have HLA-DRB1*0401 or related alleles, and increased outer surface protein A (OspA) reactivity (Kalish et al., 1993; Steere et al., 1994; Steere, 2001). A recent controlled trial of antimicrobial treatment of patients with persistent symptoms following well-documented, previously treated LB found no difference in outcome between patients who received antibiotics and those who received a placebo (Klempner et al., 2001). The study evaluated the efficacy of treatment in both seropositive and seronegative patients. Patients received either intravenous ceftriaxone at 2 g daily for 30 days, followed by oral doxycycline at 200 mg daily for 60 days, or matching intravenous and oral placebos. The cause of the patients’ symptoms was unknown; however, the evidence suggested that it was not persistent infection, and argued against the validity of a condition described as ‘chronic (refractory) Lyme disease’. The occurrence of continuing symptoms following appropriate treatment for LB has several possible explanations other than persistent infection, including initial misdiagnosis, permanent tissue damage not responsive to antibiotics, sterile inflammation in response to poorly degraded antigens, autoimmunity and triggering of conditions such as chronic fatigue syndrome or fibromyalgia (Sigal, 1994). Although it is possible that persons can acquire a degree of natural immunity against B. burgdorferi s.l. infection, cases of reinfection have been well documented (Stobierski et al., 1994), and repeated infection is probably common in highly endemic areas (Lastavica et al., 1989). It is not known whether acquired or inherited factors can lead to an increased susceptibility to infection. Studies of immunocompromised patients with early localized infection indicated that they were more likely than immunocompetent persons to develop early disseminated infection before treatment and to require more than one course of antibiotics, although a favourable outcome was noted after 1 year in both groups (Marasapin et al., 1999). The complications of untreated LB can be severe and disabling, but reports of LB as an immediate cause of death are rare and mostly unsubstantiated; in one fatal case associated with culture-confirmed Lyme carditis, the patient had concomitant babesiosis (Marcus et al., 1985).

Distribution of Lyme Borreliosis Global patterns The distribution of LB closely matches the worldwide distribution of ticks of the I. ricinus complex (see Fig. 4.1, Eisen and Lane, Chapter 4) (Schmid, 1985;

Epidemiology of Lyme Borreliosis

257

Korenberg et al., 1993; Dennis et al., 1998; Li et al., 1998). LB is endemic in certain temperate areas of the USA and Canada; it is not known to occur in Mexico. In Europe, the disease occurs in localized areas of the British Isles and throughout continental Europe, with relatively high frequency in southern Scandinavia, The Netherlands, parts of Germany and in eastern European states, such as Austria and Slovenia. Endemic levels are relatively low in France, Switzerland, Italy, Spain and Portugal, and the Balkan states. The rates of reported cases decrease from south to north in Scandinavia, and from north to south in Italy (O’Connell et al., 1998). Among states of the former USSR, it occurs from Baltic countries eastward through northern forest areas of Russia to the Pacific Coast. In China, the disease is most highly endemic in forested northeastern regions, and occurs only sporadically in widely scattered foci elsewhere (Wanchun et al., 1998). In Japan, the disease is rare, and most cases arise from exposures in limited subalpine forest areas of Nagano Prefecture of central Japan and on Hokkaido Island in northern Japan (Yanagihara and Masuzawa, 1997). The risk to humans from enzootic cycles of infection in Korea and Taiwan is unknown. Although cases of suspected LB have been reported from sub-Saharan Africa, Latin America and Australia, these are unsubstantiated: epidemiological patterns of these cases have not been defined, transmission cycles have not been described and individual cases have not been confirmed by bacterial isolation or molecular characterization of the agent. Established populations of I. ricinus do occur in some mountainous areas of North Africa and the Near East, but neither human cases nor established enzootic cycles of B. burgdorferi s.l. strains known to infect humans have been confirmed there.

Focal patterns and ecological correlates in North America The distribution of LB is focal within endemic regions and geopolitical areas such as provinces, states, counties, districts, townships and even within neighbourhoods. In the USA, cases occur mostly among persons living in coastal northeastern and mid-Atlantic regions, from Maine to Maryland, and the northcentral states of Wisconsin and Minnesota (Orloski et al., 2000; Centers for Disease Control and Prevention, 2001). High-risk areas are characterized by a combination of forest and forest-edge habitats that support white-tailed deer (Odocoileus virginianus), humid, temperate microclimatic ground-level conditions that favour Ixodes spp. ticks in all stages of their development (conditions most often provided by shade and leaf litter), and habitats that are suitable for rodent reservoirs such as P. leucopus and Tamius (chipmunk) spp. (Spielman et al., 1985; Fish, 1995; Kitron and Kazmierczak, 1997; Wilson, 1998). Much smaller numbers of LB cases occur at Pacific coastal sites where the human-biting tick I. pacificus is a bridge vector from a natural cycle of B. burgdorferi s.s. involving woodrats, mice and I. spinipalis ticks, and where lizards serve as preferred hosts of subadult ticks (Lane et al., 1991). The favoured habitat in California is open oak and bay tree woodland with dense underlying leaf litter. The black-tailed

258

D.T. Dennis and E.B. Hayes

deer (Odocoileus hemionus columbianus) is an important maintenance host for adult I. pacificus ticks. In eastern Canada, LB is a known enzoosis only in southeastern Ontario Province where there are established populations of I. scapularis ticks, although these ticks have been collected in scattered foci elsewhere in eastern maritime Canada (Canada Health and Welfare, 1998). In western Canada, scarce I. pacificus infected with B. burgdorferi s.s. have been reported from British Columbia, as have rare human cases (Canada Health and Welfare, 1993).

Focal patterns and ecological correlates in Europe The distribution of LB in Europe and Asia is also discrete. In Europe, risk is associated with residence in rural areas, but not those characterized by intensive agriculture, with recreational and leisure activities in forested natural areas set aside for public use, and with occupational activities in forested areas (O’Connell et al., 1998; Gray, 1999). Urban parks may also pose a significant risk (Guy and Farquhar, 1991; Matuschka et al., 1997; Junttila et al., 1999). The tick vectors of LB in Eurasia parasitize a great number and diversity of vertebrate hosts, and the various genospecies of B. burgdorferi s.l. are associated with particular groups of competent reservoir hosts, such as B. garinii and B. valaisiana with birds, B. afzelii with small mammals (such as various mice, voles, rats and shrews), and B. burgdorferi s.s. and B. afzelii with squirrels (Gern and Humair, 1998). Ixodes uriae, a tick that feeds on seabirds, maintains transmission cycles of B. garinii, and infected I. uriae ticks have been reported in both northern and southern hemispheres, suggesting global avian dispersal of infection (Olsen et al., 1995). Passerine birds have been suggested as sources of dispersal of infection in both North America and in Europe (Anderson et al., 1986). The usual maintenance hosts for I. ricinus and I. persulcatus ticks are ungulates, such as sheep and deer, especially the roe deer, red deer and sika deer, and various other large and medium-sized vertebrates (Gern et al., 1998). In some restricted urban parks, however, robust enzootic cycles are maintained in the absence of these animals (Juntilla et al., 1999).

Descriptive Population Data on Lyme Borreliosis in North America Emergence of Lyme borreliosis in the USA In the mid-1970s, Steere and co-workers described an outbreak of arthritis and EM in the town of Lyme and adjacent Connecticut communities (Steere et al., 1977). They named the disease after the town of its discovery, determined that the disease first appeared 3–4 years earlier, and that its period prevalence was 4.3 per 1000 persons in the study communities. Other studies identified similar small communities with rapidly emerging LB (Hanrahan et al., 1984; Steere et

Epidemiology of Lyme Borreliosis

259

al., 1986; Lastavica et al., 1989; Alpert et al., 1992). A review of reported cases in Connecticut for the period 1977–1985 suggested a threefold to eightfold increase in incidence in the communities where LB was first described, and an extension of disease from one county to all counties in Connecticut (Petersen et al., 1989). These patterns of emergence continue unabated in this state (Connecticut Department of Public Health, 2001). Rapid emergence has also been well documented in New York State. Prior to 1979, vector ticks were known only from eastern Long Island, but by 1989 they were recognized in 22 New York counties, and enzootic cycles of B. burgdorferi s.s. were described in eight adjacent counties in southern New York (White et al., 1991). Since then, spread has continued in New York mostly as a contiguous extension of established foci (Glavanakov et al., 2001). In the mid-western USA, extension of LB follows riparian fringe habitat (Kitron and Kazmierczak, 1997; Guerra et al., 2001). LB in the far-western USA is considered a relatively low risk and appears to be stable. The highest reported incidence occurs in several coastal counties of northern California. An epidemiological investigation of 230 residents of a semirural subdivision in Sonoma County, considered to one of the most endemic areas in California, found that only 1.4% of persons had serological evidence of infection with B. burgdorferi s.s., although 12.6% of residents gave a history of illness thought by them to be LB (Fritz et al., 1997). Active surveillance for LB in several of the most highly endemic counties in California, using physiciandiagnosed EM as the case criterion, yielded an annual incidence of only 5.5 per 100,000 population, a case rate substantially lower than in the northeastern USA. The authors concluded that the disease was being overdiagnosed in the area (Ley et al., 1994).

National surveillance for Lyme borreliosis in the USA In the USA, an informal system of LB reporting was established in the early 1980s, and was standardized after a uniform case definition was adopted for national surveillance, beginning in 1991 (Centers for Disease Control and Prevention, 1998). This definition consists of clinical criteria for early localized borreliosis (EM) and for disseminated borreliosis, and employs serological testing to substantiate cases presenting with signs other than EM (Table 10.1). State and local health authorities employ various procedures for passive and active surveillance in their jurisdictions, and surveillance, although mandatory in all states, is carried out with varying intensity from one place to the next. Increasingly, states have relied on laboratory-based surveillance to identify suspected cases, which then undergo various validation processes before being reported as confirmed cases. In Canada, a modified version of the USA clinical case definition was established for surveillance purposes in 1991. Clinical case reporting is mandatory in Ontario and British Columbia Provinces, the only Canadian provinces with known established enzootic cycles of B. burgdorferi s.s., and public health laboratory-based reporting occurs throughout Canada.

260

D.T. Dennis and E.B. Hayes

Table 10.1. Case definition of Lyme borreliosis for national surveillance in the USA. (Adapted from recommendations made by the Centers for Disease Control and Prevention (CDC, 1997), and modified by Steere (2001).) Skin

Erythema migrans observed by a physician. This skin lesion expands slowly over a period of days or weeks to form a large, round lesion, often with central clearing. To be counted for surveillance purposes, a solitary lesion must reach a size of at least 5 cm.

Nervous system

Lymphocytic meningitis, cranial neuritis, radiculoneuropathy or, rarely, encephalomyelitis, alone or in combination. For encephalomyelitis to be counted for surveillance purposes, there must be evidence in cerebrospinal fluid of the intrathecal production of antibodies against Borrelia burgdorferi.

Cardiovascular system

Acute-onset, high-grade (second or third degree) atrioventricular conduction defects that resolve in days or weeks and are sometimes associated with myocarditis.

Musculoskeletal system

Recurrent, brief attacks (lasting weeks to months) of objectively confirmed joint swelling in one or a few joints, sometimes followed by chronic arthritis in one or a few joints.

Laboratory evidence

Isolation of B. burgdorferi from tissue or body fluid or detection of diagnostic levels of antibody against the spirochete by the two-test approach of ELISA and Western blotting, interpreted according to the criteria of the Centers for Disease Control and Prevention and the Association of State and Territorial Public Health Laboratory Directors.a

a

In a person with acute disease of less than 1 month’s duration, IgM and IgG antibody responses should be measured in serum samples obtained during the acute and convalescent phases. A Western blot for IgM antibodies is considered positive if at least two of the following three bands are present: 23, 39 and 41 kDa. A blot IgG antibodies is considered positive if at least five of the following ten bands are present: 18, 23, 28, 30, 39, 41, 45, 58, 66 and 93 kDa. Only the IgG response should be used to support the diagnosis after the first month of infection; after that time, an IgM response alone is likely to represent a false positive result.

As is the case with other infectious diseases, LB reporting in the USA is unenforced, and poorly and irregularly complied with. In 1991–1992, in Connecticut, more than 80% of all cases of LB registered with the state were reported by physicians from one of four primary care specialities, even though only 7% of

Epidemiology of Lyme Borreliosis

261

physicians in these specialities had reported cases (Centers for Disease Control and Prevention, 1993). Follow-up interviews with a sample of these physicians suggested that at best only 16% of all cases diagnosed and treated as incident cases of LB had been reported (Meek et al., 1996). A study in Maryland suggested that the disease was under-reported by 10- to 12-fold in that state, and that many more patients were seen and treated for presumptive LB and for tickbite alone than patients meeting the case criteria for reporting (Coyle et al., 1996). Overdiagnosis and misclassification of cases are also important surveillance problems, especially in areas of low or questionable endemicity (Steere et al., 1993; Seltzer and Shapiro, 1996). Too much reliance is placed on serological tests for making the diagnosis of LB, and testing is often applied inappropriately, especially in persons with suspected EM, and in persons with vague symptoms. ELISA tests are, however, generally highly sensitive in early disseminated and late LB, and they can be helpful in these circumstances in ruling out the diagnosis. The two-tiered approach of an ELISA followed by immunoblotting of ELISA-positive sera is highly specific, and can be usefully applied to ‘rule in’ infection for individual case evaluations and for seroepidemiological studies (Fritz et al., 1997; Trevejo et al., 2001). In the USA, LB accounts for more than 90% of all reports of vector-borne infectious disease. In 1999, 16,273 cases were reported by 45 states (overall incidence of 6.0 per 100,000 population) (Centers for Disease Control and Prevention, 2001). This compares with 497 cases by 11 states in 1982, and continues a secular trend of rising case numbers (Fig. 10.1). The map of the distribution of cases by county clearly shows geographical concentrations in the northeastern, mid-Atlantic and north-central regions, and in northern

18 16

Cases (thousands)

14 12 10 8 6 4 2 0 1982

1984

1986

1988

1990

1992

1994

1996

1998

Fig. 10.1. Reported cases of Lyme borreliosis in the USA, 1982–2000.

2000

262

D.T. Dennis and E.B. Hayes

California (Fig. 10.2). Nine states reporting incidences higher than the national rate (i.e. Connecticut, 98.0; Rhode Island, 55.1; New York, 24.2; Pennsylvania, 23.2; Delaware, 22.2; New Jersey, 21.1; Maryland, 17.4; Massachusetts, 12.7; Wisconsin, 9.3 per 100,000 population) accounted for 92.0% of the nationally reported cases. Although cases are regularly reported by about 45 states, reporting is by residence of case, not by place of exposure. Epidemiological data suggest that only about 15 states are endemic for LB, with sporadic cases occurring in some bordering states, and evidence is lacking for transmission of infection in about 25 other states. Cases reported from the latter states are likely to have had exposures elsewhere or to represent case misclassifications. In 1999, county of residence was available for 16,214 (99.6%) nationally reported cases. Among the 3143 USA counties, 713 (22.7%) had at least one case reported during 1999; 90% of these cases were from just 109 (15.3%) reporting counties in ten states. The highest county-specific incidence (950.7 per 100,000) occurred in Nantucket County, which comprises a small population residing on the offshore Massachusetts island of Nantucket. Incidence exceeded 100 cases per 100,000 persons in 24 counties. Several counties in northeastern states routinely report annual incidence rates of 400 per 100,000 population or greater. LB affects persons in all age groups, but the rate of disease incidence is bimodal with the highest rates found in children less than 15 years of age and in adults 35 years of age and older. Consistently, over time and from one geographical area to another, surveillance statistics show a pronounced downturn in the case rates in the age group 15–35 years, with least rates in the third decade of life (Fig. 10.3). In the USA from 1992–1998, the median age of reported cases was 30 years (Centers for Disease Control and Prevention, 2001). Overall, slightly

Fig. 10.2. Reported cases of Lyme borreliosis in the USA by county, in 2000.

Epidemiology of Lyme Borreliosis

263

8

Male Female

7

Cases per 100,000

6 5 4 3 2 1 0 0-4

5-9 10-14 15-19 20-24 25-29 30-34 35-39 40-44 45-49 50-54 55-59 60-64 65-69 70-74 75-79 80-84 > 84

Age (years)

Fig. 10.3. Mean annual incidence of reported cases of Lyme borreliosis by age group and sex in the USA, 1992–2000.

more males than females were reported as cases (51.9%); the difference was found to be greatest in children aged 5–14 years, and in adults greater than 60 years of age. The majority (about 70%) of cases experience disease onset in June, July or August, although cases occur in all months of the year (Fig. 10.4). Statistics on clinical presentations of reported cases are not available nationally, but are recorded by some states. Connecticut, the state with the highest reported case 30,000

25,000

Cases

20,000

15,000

10,000

5000

0 Jan

Feb

Mar

Apr

May

Jun

Jul

Aug

Sep

Oct

Nov

Reported month of disease onset

Fig. 10.4. Month of Lyme borreliosis onset for reported cases in the USA, 1992–2000.

Dec

264

D.T. Dennis and E.B. Hayes

rates and the most complete surveillance analyses, recorded nearly 4000 cases in 2000; 65% were reports of cases with EM alone, 7% had EM and a systemic manifestation of LB, and 28% had one or more systemic manifestations and a positive serological test for antibodies to B. burgdorferi s.s. Of the more than 1000 disseminated LB cases, arthritis was reported in 69%, neurological manifestations in 36% and cardiac complications in only 2% of cases (Connecticut Department of Public Health, 2001). Published clinical case series, which represent a different patient population from nationally reported cases, describe higher proportions of patients with EM rashes (Nadelman and Wormser, 1995).

Descriptive Population Data on Lyme Borreliosis in Europe Surveillance statistics in Europe are based on varying case criteria and on disparate systems of data collection. LB is nationally reportable only in Scotland, Slovenia and Denmark (neuroborreliosis); however, a number of jurisdictions below the state level have developed surveillance systems that allow estimations of incidence and its trends. Nearly all countries have made efforts to determine the national scope and magnitude of the problem, including the total number of new cases annually, and the mapping of the cases by place of residence or report. A report of a 1995 World Health Organization workshop on LB diagnosis and surveillance provided estimates of the occurrence of LB in several European countries. Estimates of annual incidence ranged from 0.3 and 0.6 per 100,000 in the UK and Ireland, respectively, 16–39 in several central European countries, 55 in Bulgaria, and 120 and 130 per 100,000 in Slovenia and Austria, respectively. Annual numbers of cases ranged from 30 in Ireland to 20,000 in Germany, with a total estimate of 60,000 for the ten reporting countries (O’Connell et al., 1998). The existence of several pathogenic B. burgdorferi genospecies complicates surveillance. These organisms are not evenly distributed geographically and they demonstrate differing tick and vertebrate host preferences; serological testing needs to consider different antigenic expressions of the various strains, especially the variability in expression of OspC, and clinical manifestations differ somewhat depending on which genospecies is responsible for infection (van Dam et al., 1993; Wang et al., 1999). In general, B. afzelii infection is associated more with cutaneous disease (especially ACA), B. garinii more with neurological manifestations and B. burgdorferi s.s. more with arthritis (Wang et al., 1999). The case definition for LB developed by the European Union Concerted Action on Risk Assessment in Lyme Borreliosis (EUCALB) is more clinically detailed than that used in the USA, and incorporates criteria for lymphocytoma and ACA (Stanek et al., 1996), which are very rare manifestations of infection in North America. To support the clinical diagnosis, an attempt has been made by EUCALB to develop standardized criteria for the interpretation of

Epidemiology of Lyme Borreliosis

265

immunoblots, but internationally useful criteria have been hard to achieve (Robertson et al., 2000b). Investigators in Sweden conducted a prospective population-based survey of cases in seven southern counties in a 1 year period, 1992–1993 (Berglund et al., 1995). The study area was considered the most highly endemic region in Sweden and comprised 24% (2.13 million) of the population. Primary care physicians were asked to report all cases of illness with diagnosed EM or one or more objective clinical manifestations of disseminated disease plus serological confirmation. The study identified 1471 cases giving an overall annual incidence of 69 per 100,000 population (range 26–160). EM was seen in 77% of cases, followed by neuroborreliosis (16%) and arthritis (7%). Carditis was extremely rare (only two cases). The incidence varied according to area, and was highest in the eastern coastal area. Cases occurred in all months of the year, but tick bites were most frequent in July, and onsets of EM peaked in August. As in the USA, the incidence by age group was bimodal, with highest rates in children in the age group 5–9 years and in adults aged 60–75 years. Nearly 80% of patients remembered an antecedent tick bite; bites on the head and neck region were more common in children than adults, and were associated with an increased risk of neuroborreliosis. The epidemiological picture in southern Sweden was very similar to that seen in endemic areas of North America. In a retrospective study in The Netherlands in 1995 (de Mik et al., 1997), all general practitioners in the country were asked to provide the number of patients in their practice who had been seen in the previous year for tick bite or for EM rash illness. The response rate was 80% and using extrapolations it was estimated that the practitioners saw approximately 33,000 patients with tick bites and 6500 patients with EM (approximately 43 per 100,000 population). Practices were geocoded by local government area (n = 633) and linked to geographical information system data on several major ecological variables. Ecological factors positively associated with both tick bites and EM were the proportion of the area covered by woods, sandy soil, dry uncultivated land, the number of tourist-nights per inhabitant and sheep population density. Despite the limitations of a retrospective, non-validated survey, this innovative study provided a quick estimate of approximate national rates of incident disease and some ecological correlates of risk important for developing educative messages. Slovenia health authorities have required reporting of LB since 1988. A review of the epidemiological situation there noted that the annual incidence of reported cases has, since 1994, regularly exceeded 100 per 100,000 population overall, and is more than 200 per 100,000 in some parts of the country (Strle, 1999). Under-reporting is an acknowledged problem. Since 1994, the northeastern Brandenburg region of Germany, which surrounds Berlin, has had an established reporting system using uniform case definitions for early localized and disseminated LB. The incidence of reported cases has increased over time, but it is unclear how much of this increase may be due to reporting artefact such as increased recognition and compliance. The annual incidence in 1997 was as high as 48 per 100,000 in some districts (Talaska, 1998).

266

D.T. Dennis and E.B. Hayes

To assess the incidence of LB in central Europe, a 12 month (1996–1997) population-based surveillance study was conducted in a population of 279,000 persons in the Wurzburg region of Germany (Huppertz et al., 1999). All physicians were requested to report cases meeting a case definition developed for the study. This definition required the presence of an EM rash of 5 cm in diameter or greater, lymphocytoma, or an objective manifestation of later disseminated infection with serological confirmation. A total of 313 cases (111 per 100,000) were reported. The highest rates occurred in children and elderly adults living in wooded as opposed to agricultural or urban areas. EM was the only manifestation in 89% of patients. In a prospective serological survey in southeastern Bavaria, Germany, nearly 5000 adults having no history of LB were enrolled (Reimer et al., 1999). Blood samples obtained at times 0, 1 year and 2 years were tested using IgG ELISA and substantiated by Western immunoblotting. In the follow-up period, which included more than 55,000 person-observation months, 33 incident cases of LB were diagnosed, giving an incidence of 7.1 per 1000 person-observation years; additionally, the authors reported an asymptomatic seroconversion of 7.6 per 1000 person-observation years. On this basis, it was concluded that the incidence of infection in this poorly defined sample was 1.5% per year, and the incidence of diagnosed LB was 0.7% per year. Several seroprevalence studies for LB in selected European population groups considered to be at high risk have been reported, with rates varying from 5 to 20%, and exceeding by twofold or more the rates of reported LB in these participants (O’Connell et al., 1998). The largest number of studies have examined populations thought to be at high risk because of recreational or occupational exposures to infective ticks, including orienteers, foresters, military personnel, farmers and other classes of outdoor workers (O’Connell et al., 1998). Although these studies reported elevated levels of seropositivity in the at-risk groups, and increasing levels with a longer duration of exposure, the pretest probabilities of LB in the study and comparison populations were generally poorly defined, the serodiagnostic methods used were unstandardized and the confidence intervals around the prevalence estimates were not given. In a sero-epidemiological study of 950 Swiss orienteers, diagnostic levels of IgG ELISA antibodies against B31 strain whole-cell sonicated antigens were detected in 26% of bloods taken in the spring of the year from persons in the study group, and in 3.9% and 6.0% of two comparison groups (Fahrer et al., 1991). Using questionnaire responses, only 1.9% and 3.1% of the study group had a past history compatible with confirmed or probable LB. Seroconversions occurred in 45 (8.1%) of 755 orienteers between the spring bleed and a second bleed 6 months later; however, questionnaire data indicated that only one (2.2%) of those who seroconverted developed clinically recognizable LB. A study in Austria showed that the seasonal peak of EM cases occurred in July, with the peak of neuroborreliosis (NB) cases occurring approximately 1 month later (Fig. 10.5). This is very similar to the seasonal pattern of disease onset in the USA (Fig. 10.4), despite differences in the life cycles of I. ricinus in

Epidemiology of Lyme Borreliosis

267

Europe and I. scapularis in the USA (Gray, 1998), though in both regions the nymphal stage is thought to be responsible for most cases (Falco et al., 1999; Robertson et al., 2000a).

Determinants of Infection Risk factors for Lyme borreliosis In the USA, the principal risk factor for LB is permanent or seasonal residence in an area with high infestation by infected ticks (Spielman, 1994; Fish, 1995; Cromley et al., 1998). Exposures to infected ticks are most likely to occur in periresidential areas that are relatively undisturbed, i.e. wooded tracts, and tall grass and brushy scrub at edges of gardens (yards), and are considerably less likely to occur on lawns and other maintained areas (Maupin et al., 1991). A case–control study was conducted in Hunterdon County, New Jersey, to determine risk factors for LB in reported cases. Comparisons of age-matched cases and controls showed that rural residence, clearing of peri-residential brush during spring and summer months, and the presence of stone walls, woods, deer or a bird feeder on residential property were each associated with incident disease (Orloski et al., 1998). A study of I. scapularis-borne illness (LB and ehrlichiosis) in Wisconsin compared exposure factors in cases and their matched controls (Belongia et al., 1999). 50,0

EM

Seasonal occurrence (%)

40,0

NB

30,0

20,0

10,0

0,0 J

F

M

A

M

J

J

A

S

O

N

D

Fig. 10.5. Comparison of seasonal occurrence of erythema migrans (EM, n = 566) and neuroborreliosis (NB, n = 148) in northeastern Austria, 1994–1995 (G. Stanek et al., unpublished results).

268

D.T. Dennis and E.B. Hayes

Results of multivariate logistic regression modelling identified residence in a rural neighbourhood and a history of camping as risk factors for illness. Using conditional logistic regression, recognized tick bite in the month before symptom onset, jogging, clearing brush off property, camping and residential property size > 2 acres were significantly associated with tick-borne illness. The study was unable to document the effectiveness of insect repellent use or other personal protective measures. In California, a case–control study of persons with physician-diagnosed EM and matched controls found that the observation of deer and lizards around the home, a recognized exposure to ticks, and outdoor leisure activities were risk factors there for LB (Ley et al., 1995). In the original epidemiological studies of Lyme arthritis in coastal Connecticut, USA, ownership of cats was found to be weakly associated with an increased risk of acquiring LB (Steere et al., 1978), and domestic cats have been observed to bring unattached nymphal ticks into homes (Curran and Fish, 1989). The association of cat ownership and LB in their owners has not been substantiated in other studies. A study of dogs and persons living in the same households in two highly endemic areas in Massachusetts showed that dogs were more likely to have serological evidence of B. burgdorferi s.s. infection than their human coresidents (Lindenmayer et al., 1991), and that dog ownership was not associated with an increased risk to their owners (Eng et al., 1988). Canine serosurveys have been useful in identifying geographical areas of risk (Daniels et al., 1993; Guerra et al., 2001). A number of studies have shown increased risk among certain occupational groups. A study of employees in New Jersey showed that outdoor workers were considerably more likely to have had LB than indoor workers (Bowen et al., 1984; Schwartz and Goldstein, 1990). In New York, workers in endemic counties with a history of outdoor employment were twice as likely to be seropositive as persons without such a history. Although the difference was not statistically significant, the seroprevalence rate of outdoor employees was 5.9 times higher than a comparison group of anonymous blood donors. The most important risk in New York workers was a history of spending 30 or more hours of leisure a week outdoors (Smith et al., 1988). Studies in Europe and Asia have shown that foresters are at high risk of infection: serological evidence of B. burgdorferi s.l. exposure in forest workers has been positively associated with workers’ age, history of tick bite and history of EM (Kuiper et al., 1991; Ai et al., 1994; Nakama et al., 1994; Rath et al., 1996).

Risk for adverse outcomes of pregnancy Several epidemiological studies have examined whether B. burgdorferi s.l. infection during pregnancy results in adverse outcomes to the fetus, including stillbirth, abortion and various congenital defects. A prospective study of 2000 prenatal patients in a highly endemic area of New York State concluded that maternal

Epidemiology of Lyme Borreliosis

269

LB before conception and during pregnancy was not associated with fetal death, birth weight, length of gestation at delivery or congenital malformations (Strobino et al., 1993). A retrospective case–control study in the same population determined that women who had been bitten by a tick or treated for LB during or before pregnancy were not at an increased risk for giving birth to a child with a congenital heart defect (Strobino et al., 1999). Cardiac defects were selected as the outcome measure because they comprise a large proportion of congenital defects in general, and because previous observational studies had suggested a possible association with maternal LB. A large population-based study using an ecological design showed no positive associations between adverse outcomes of pregnancy and endemic levels of LB in the county of residence of the mothers (Centers for Disease Control and Prevention, unpublished data).

Cost-effectiveness Analyses Decision trees have been used to evaluate the cost per case averted (cost-effectiveness) of vaccinating persons against LB. One study, assuming a 0.80 probability of diagnosing and treating early LB, a 0.005 probability of contracting LB and a vaccination cost of $50 per year, yielded a cost to society of $4466 per case averted. Increasing the probability of contracting LB to 0.03 and the cost of vaccination to $100 per year resulted in a mean net societal savings per case averted of $3377 (Meltzer et al., 1999). Of the variables examined, the incidence of LB in the population at risk had the greatest impact on the cost-effectiveness of vaccination. The likelihood of early diagnosis and treatment also had a substantial impact on the model because of the potential for reduced incidence of complications of infection. In a second study, a decision-analytic model was developed that measured cost-effectiveness in terms of quality-adjusted life years (QALY) (Shadick et al., 2001). Vaccinating 10,000 residents in endemic areas with an expected annual incidence of 0.01 averted 202 cases of LB during a 10 year period. The additional cost per QALY gained compared with no vaccination was $62,300. If a yearly booster shot is required for persisting efficacy, the marginal cost-effectiveness ratio increased to $72,700 per QALY. The authors of both studies concluded that vaccination was economically attractive only when the annual incidence of infection was greater than 0.01. These cost-effectiveness estimates were important in the development of national recommendations for vaccine use based on expected risk of exposure to infective ticks, advising that vaccination be considered only for persons who reside, work or take recreation in areas of high or moderate risk and who also engage in activities that result in exposure to tick-infested habitats. The cost-effectiveness for treating tick bite with 2 weeks of oral doxycyline was examined by Magid et al. (1992). The authors concluded that empirical treatment is indicated when the probability of B. burgdorferi s.l. infection after a bite is 0.036 or higher, and that this treatment may be preferred when the probability ranges between 0.01 and 0.035, since at these levels the value of

270

D.T. Dennis and E.B. Hayes

preventing major complications of illness outweigh minor complications of treatment; however, at probabilities less than 0.01, the model predicted that the costs of the excess minor complications incurred would exceed the costs of major disease complications averted. A study of the costs of serological diagnostic testing for LB in Maryland, an endemic state in the mid-Atlantic region of the USA having a population of about 5 million persons, showed that from 1992 to 1995 approximately 30,000 tests were performed annually at a cost to Maryland residents of about $2 million in direct medical costs per year (Strickland et al., 1997). The percentage of positive ELISA and Western immunoblot tests did not exceed 7.9% for any one year, emphasizing the misuse of tests for screening and other inappropriate purposes and the risk of misleading test results, especially of ELISA results. In a separate study, it was determined that more patients were seen and treated for tick bites and for presumptive LB than for substantiated cases (Coyle et al., 1996). A diagnosis of LB is best made by evaluating the patient’s clinical findings, combined with an epidemiological assessment of exposure and supported by serological testing (Tugwell et al., 1997). A two-tier serodiagnostic approach, using the ELISA as a first test, followed by immunoblotting of sera found to be positive or equivocal, is recommended by authorities both in the USA (Centers for Disease Control and Prevention, 1995) and in Europe (Robertson et al., 2000b) as the strategy providing the greatest accuracy and reliability. Referring to this algorithm, the American College of Physicians concluded that testing for LB was in general not useful if the pretest probabilities were less than 0.20 or greater than 0.80 (American College of Physicians, 1997; Tugwell et al., 1997). A costeffectiveness analysis of differing test-treatment strategies concluded that the most economically attractive strategy for persons with myalgia alone was no testing and no treatment; empirical treatment was most attractive for patients with EM, and two-tier testing was most attractive for persons with oligoarticular arthritis with one or no other features suggestive of LB (Nichol et al., 1998). These findings complement earlier cost-effectiveness studies by Lightfoot et al. (1993), who determined that empirical treatment with intravenous antibiotics of seropositive patients who have only non-specific symptoms incurs costs and risks that far exceed any benefits of such treatment. A follow-up cost-effectiveness study by Eckman et al. (1997) used a simulation model to analyse decision trees and perform sensitivity analyses in a consideration of oral compared with intravenous antibiotic therapy of two groups of patients: those with early LB and those with Lyme arthritis who had no clinical evidence of neurological involvement. These analyses showed that, for both groups of patients, treatment with oral doxycyline was as effective as intravenous ceftriaxone and was less costly.

Prevention of Lyme Borreliosis Infection with B. burgdorferi s.l. can be prevented by reducing exposure to infective ticks or by reducing the likelihood that ticks will transmit infection when

Epidemiology of Lyme Borreliosis

271

exposure is unavoidable. Strategies to reduce tick densities in the environment and to reduce infection rates in ticks are discussed by Stafford and Kiken in Chapter 12. The use of repellents, personally applied acaricides and clothing barriers to tick attachment, as well as the prompt removal of ticks once they are attached, should theoretically reduce the risk of acquiring LB, but studies to evaluate the efficacy of such methods have not provided conclusive evidence of their effectiveness (Smith et al., 1988; Ley et al., 1995; Belongia et al., 1999; Orloski et al., 2000). A vaccine recently developed for use in the USA was nearly 80% efficacious in preventing LB (Steere et al., 1998) and recommendations for its use were published (Centers for Disease Control and Prevention, 1999). The vaccine and its use are discussed in detail by Hayes and Schriefer in Chapter 11. Using a computer simulation model and basic epidemiological concepts, Hayes et al. (1999) estimated the effectiveness of several interventions for preventing LB in a hypothetical endemic community where the disease is transmitted by I. scapularis ticks. The model examined such interventions as the spreading of granular acaricides, topical application of acaricides to deer, reduction in deer density, deer exclusion, reduction of tick habitat through landscape modification, education to increase the use of personal protective measures and administration of vaccine. The effectiveness of these interventions was measured by numbers of LB cases prevented. The model highlighted the concept that the effectiveness of any intervention is determined by its theoretical efficacy and by the level of engagement of the target population. Populations living in LB-endemic areas may be exposed to large numbers of bites by vector ticks (Campbell et al., 1998). Several early studies, including a meta-analysis, did not show that treatment after I. scapularis tick bites is efficacious in preventing LB (Warshafsky et al., 1996). Recently, Nadelman et al. (2001) conducted a large, randomized, double-blind, placebo-controlled trial in an area of New York where LB is endemic, and reported that a single 200 mg dose of doxycyline given within 72 h after an I. scapularis tick bite prevented the development of the disease. The ease of administration, safety and low cost of this regimen may justify its use as an adjunct to other prevention practices in selected high-risk circumstances. The development and implementation of effective prevention measures requires an in-depth understanding of LB epidemiology, which is in turn dependent on knowledge of the complex biology of the pathogens, B. burgdorferi s.l., and of their interface with humans. The increasingly evident genetic diversity of members of this species complex, and their ecological adaptations and changing human–parasite interactions, has implications for pathogenicity, laboratory diagnosis, exposure risks, and for the development of effective vaccines and other methods of prevention and control. This great breadth of diversity presents a formidable public health challenge in both Old and New World circumstances.

272

D.T. Dennis and E.B. Hayes

References Ai, C., Zhang, W. and Zhao, J. (1994) Seroepidemiology of Lyme disease in an endemic area in China. Microbiology and Immunology 38, 505–509. Alpert, B., Esin, J., Sivak, S.L. and Wormser, G.P. (1992) Incidence and prevalence of Lyme disease in a suburban Westchester County community. New York State Journal of Medicine 92, 5–8. American College of Physicians (1997) Clinical guideline, Part 1. Guidelines for laboratory evaluation in the diagnosis of Lyme disease. Annals of Internal Medicine 127, 1106–1108. Anderson, J.F., Johnson, R.C., Magnarelli, L.A. and Hyde, F.W. (1986) Involvement of birds in the epidemiology of the Lyme disease agent Borrelia burgdorferi. Infection and Immunity 51, 394–396. Barbour, A.G. (1996) Does Lyme disease occur in the south? A survey of emerging tick-borne infections in the region. American Journal of Medical Science 311, 34–40. Barbour, A.G., Maupin, G.O., Teltow, G.J., Carter, C.J. and Piesman, J. (1996) Identification of an uncultivable Borrelia species in the hard tick Ambloymma americanum: possible agent of a Lyme diseaselike illness. Journal of Infectious Diseases 173, 403–409. Belongia, E.A., Reed, K.D., Mitchell, P.D., Chyou, P.H., Mueller-Rizner, N., Finkel, M.F. and Schriefer, M.E. (1999) Clinical and epidemiological features of early Lyme disease and human granulocytic ehrlichiosis in Wisconsin. Clinical Infectious Diseases 29, 1472–1477. Berglund, J., Eitrem, R., Ornstein, K., Lindberg, A., Ringner, A., Elmrud, H., Carlsson, M., Runehagen, A., Svanborg, C. and Norrby, R. (1995) An epidemiologic study of Lyme disease in southern Sweden. New England Journal of Medicine 333, 1319–1324. Bowen, G.S., Schulze, T.L., Hayne, C. and

Parkin, W.E. (1984) A focus of Lyme disease in Monmouth County, New Jersey. American Journal of Epidemiology 120, 387–394. Brown, R.N. and Lane, R.S. (1996) Reservoir competence of four chaparral-dwelling rodents for Borrelia burdorferi in California. American Journal of Tropical Medicine and Hygiene 54, 84–91. Campbell, G.L., Paul, W.S., Schriefer, M.E., Craven, R.B., Robbins, K.E. and Dennis, D.T. (1995) Epidemiologic and diagnostic studies of patients with suspected early Lyme disease, Missouri, 1990–1993. Journal of Infectious Diseases 172, 470–480. Campbell, G.L., Fritz, C.L., Fish, D., Nowakowski, J., Nadelman, R.B. and Wormser, G.P. (1998) Estimation of the incidence of Lyme disease. American Journal of Epidemiology 148, 1018–1026. Canada Health and Welfare (1993) Isolation of Borrelia burgdorferi in British Columbia. In: Canada Communicable Disease Report, pp. 19–24, 204–205. Canada Health and Welfare (1998) Distribution of Ixodes pacificus and Ixodes scapularis re concurrent babesiosis and Lyme disease. In: Canada Communicable Disease Report, pp. 24–15. Centers for Disease Control and Prevention (1993) Physician reporting of Lyme disease – Connecticut, 1991–1992. Morbidity and Mortality Weekly Report 42, 348–350. Centers for Disease Control and Prevention (1995) Recommendations for test performance and interpretation from the Second National Conference on Serologic Diagnosis of Lyme Disease. Mortality and Morbidity Weekly Report 44, 590–591. Centers for Disease Control and Prevention (1997) Case definitions for infectious conditions under public health surveillance. Morbidity and Mortality Weekly Report 46, 20–21.

Epidemiology of Lyme Borreliosis

Centers for Disease Control and Prevention (1998) Preventing emerging infectious diseases: a strategy for the 21st century. Morbidity and Mortality Weekly Report 47, 1–14. Centers for Disease Control and Prevention (1999) Recommendations for the use of Lyme disease vaccine: recommendations of the Advisory Committee on Immunization Practices (ACIP). Morbidity and Mortality Weekly Report 48, 1–25. Centers for Disease Control and Prevention (2001) Lyme disease – United States, 1999. Morbidity and Mortality Weekly Report 50, 181–185. Connecticut Department of Public Health (2001) Lyme disease – Connecticut, 2000. Connecticut Epidemiology 21, 9–12. Coyle, B.S., Strickland, G.T., Liang, Y.L., Pena, C., McCarter, R. and Israel, E. (1996) The public health impact of Lyme disease in Maryland. Journal of Infectious Diseases 173, 1260–1262. Cromley, E.K., Cartter, M.L., Mrozinski, R.D. and Ertel, S.H. (1998) Residential setting as a risk factor for Lyme disease in a hyperendemic region. American Journal of Epidemiology 147, 472–477. Curran, K.L. and Fish, D. (1989) Increased risk of Lyme disease for cat owners. New England Journal of Medicine 320, 183. van Dam, A.P., Kuiper, H., Vos, K., Widjojokusumo, A., de Jongh, B.M., Spanjaard, L., Ramselaar, A.C.P., Kramer, M.D. and Dankert, J. (1993) Different genospecies of Borrelia burgdorferi are associated with distinct clinical manifestations of Lyme borreliosis. Clinical Infectious Diseases 17, 708–717. Daniels, T.J., Fish, D., Levine, J.F., Greco, M.A., Eaton, A.T., Padgett, P.J. and LaPointe, D.A. (1993) Canine exposure to Borrelia burgdorferi and prevalence of Ixodes dammini (Acari: Ixodidae) on deer as a measure of Lyme disease risk in the northeastern United States. Journal of Medical Entomology 30, 171–178. Dattwyler, R.J., Volkman, D.J., Luft, B.J.,

273

Halperin, J.J., Thomas, J. and Golightly, M.G. (1988) Seronegative Lyme disease: dissociation of specific Tand B-lymphocyte to Borrelia burgdorferi. New England Journal of Medicine 319, 1441–1446. Dennis, D.T., Nekomoto, T.S., Victor, J.C., Paul, W.S. and Piesman, J. (1998) Reported distribution of Ixodes scapularis and Ixodes pacificus (Acari: Ixodidae) in the United States. Journal of Medical Entomology 35, 629–638. Eckman, M.H., Steere, A.C., Kalish, R.A. and Parker, S.G. (1997) Cost-effectiveness of oral as compared with intravenous antibiotic therapy for patients with early Lyme disease or Lyme arthritis. New England Journal of Medicine 337, 357–363. Eng, T.R., Wilson, M.I., Spielman, A. and Lastavica, C.C. (1988) Greater risk of Borrelia burgdorferi infection in dogs than in people. Journal of Infectious Diseases 158, 1410–1411. Fahrer, H., van der Linden, S.M., Sauvain, M.J., Gern, L., Zhioua, E. and Aeschlimann, A. (1991) The prevalence and incidence of clinical and asymptomatic Lyme borreliosis in a population at risk. Journal of Infectious Diseases 163, 305–310. Falco, R.C., McKenna, D.F., Daniels, T.J., Nadelman, R.B., Nowakowski, J., Fish, D. and Wormser, G.P. (1999) Temporal relation between Ixodes scapularis abundance and risk for Lyme disease associated with erythema migrans. American Journal of Epidemiology 149, 771–776. Felz, M.W., Chandler, F.W., Oliver, J.H., Rahn, D.W. and Schriefer, M.E. (1999) Solitary erythema migrans in Georgia and South Carolina. Archives of Dermatology 135, 1317–1326. Fish, D. (1995) Environmental risk and prevention of Lyme disease. American Journal of Medicine 98, 4A-2S-9S. Fritz, C.L., Kjemtrup, A.M., Conrad, P.A., Flores, G.R., Campbell, G., Schriefer, M.E., Gallo, D. and Vugia, D.J. (1997)

274

D.T. Dennis and E.B. Hayes

Seroepidemiology of emerging tickborne infectious diseases in a northern California community. Journal of Infectious Diseases 175, 1432–1439. Gern, L. and Humair, P.-F. (1998) Natural history of Borrelia burgdorferi sensu lato. Weiner Klinische Wochenschrift 110, 856–858. Gern, L., Estrada-Pena, A., Frandsen, F., Gray, J.S., Jaenson, T.G.T., Jongejan, F., Kahl, O., Korenberg, E., Mehl, R. and Nuttall, P.A. (1998) European reservoir hosts of Borrelia burgdorferi sensu lato. Zentralblatt für Bakteriologie 287, 196–204. Glavanakov, S., White, D.J., Caraco, T., Lapenis A., Robinson, G.R., Szymanski, B.K. and Maniatty, W.A. (2001) Lyme disease in New York State: spatial pattern at a regional scale. American Journal of Tropical Medicine and Hygiene 65, 538–545. Godsey, M.S., Amundson, T.E., Burgess, E.C., Schell, W., Davis, J.P., Kaslow, R. and Edelman, R. (1987) Lyme disease ecology in Wisconsin: distribution and host preferences of Ixodes dammini, and prevalence of antibody to Borrelia burgdorferi in small mammals. American Journal of Tropical Medicine and Hygiene 37, 180–187. Gorelova, N.B., Korenberg, E.I., Kovalevskii, Y.V. and Shcherbakov, S.V. (1995) Small mammals as reservoir hosts for Borrelia in Russia. Zentrablatt für Bakteriologie 282, 315–322. Gray, J.S. (1998) The ecology of ticks transmitting Lyme borreliosis. Experimental and Applied Acarology 22, 249–258. Gray, J. (1999) Risk assessment in Lyme borreliosis. Weiner Klinische Wochenschrift 111, 990–993. Guerra, M.A., Walker, E.D. and Kitron, U. (2001) Canine surveillance for Lyme borreliosis in Wisconsin and northern Illinois: geographic distribution and risk factor analysis. American Journal of Tropical Medicine and Hygiene 65, 546–552.

Guy, E.C. and Farquhar, R.G. (1991) Borrelia burgdorferi in urban parks. Lancet 338, 253. Hanrahan, J.P., Benach, J.L., Coleman, J.L., Bosler, E.M., Morse, D.L., Cameron, D.J., Edelman, R. and Kaslow, R.A. (1984) Incidence and cumulative frequency of endemic Lyme disease in a community. Journal of Infectious Diseases 150, 489–496. Hayes, E.B., Maupin, G.O., Mount, G.A. and Piesman, J. (1999) Assessing the prevention effectiveness of local Lyme disease control. Journal of Public Health Management Practice 5, 84–92. Huppertz, H.I., Bohme, M., Standaert, S.M., Karch, H. and Plotkin, S.A. (1999) Incidence of Lyme borreliosis in the Wurzburg region of Germany. European Journal of Clinical Microbiology and Infectious Diseases 18, 697–703. Institute of Medicine (1992) Infections: Microbial Threats of Health in the United States. National Academy Press, Washington, DC. James, A.M., Liveris, D., Wormser, G.P., Schwartz, I., Montecalvo, M.A. and Johnson, B.J.B. (2001) Borrelia lonstari infection after a bite by an Amblyomma americanum tick. Journal of Infectious Diseases 183, 1810–1814. Junttila, J., Peltomaa, M., Soini, H., Marjamaki, M. and Viljanen, M.K. (1999) Prevalence of Borrelia burgdorferi in Ixodes ricinus ticks in urban recreational areas of Helsinki. Journal of Clinical Microbiology 37, 1361–1365. Kahl, O., Janetzki-Mittmann, C., Gray, J.S., Jonas, R., Stein, J. and de Boer, R. (1998) Risk of infection with Borrelia burgdorferi sensu lato for a host in relation to the duration of nymphal Ixodes ricinus feeding and the method of tick removal. Zentralblatt für Bakterologie 287, 41–52. Kalish, R.A., Leong, J.M. and Steere, A.C. (1993) Association of treatment-resistant chronic Lyme arthritis with HLA-DR4 and antibody reactivity to OspA and

Epidemiology of Lyme Borreliosis

OspB of Borrelia burgdorferi. Infection and Immunity 61, 2774–2779. Kirkland, K.B., Klimko, T.B., Meriweather, R.A., Schriefer, M., Levin, M., Levine, J., MacKenzie, W.R. and Dennis, D.T. (1997) Erythema migrans-like rash illness at a camp in North Carolina. Archives of Internal Medicine 157, 2635–2641. Kitron, U. and Kazmierczak, J.J. (1997) Spatial analysis of the distribution of Lyme disease in Wisconsin. American Journal of Epidemiology 145, 558–566. Klempner, M.S., Hu, L.T., Evans, J., Schmid, C.H., Johnson, G.M., Trevino, R.P., Norton, D., Levy, L., Wall, D., McCall, J., Kosinski, M. and Weinstein, A. (2001) Two controlled trials of antibiotic treatment in patients with persistent symptoms and a history of Lyme disease. New England Journal of Medicine 345, 85–91. Korenberg, E.I. (1994) Comparative ecology and epidemiology of Lyme disease and tick-borne encephalitis in the former Soviet Union. Parasitology Today 10, 157–160. Korenberg, E.I., Kryuchechnikov, V.N. and Kovalevsky, Y.V. (1993) Advances in investigations of Lyme borreliosis in the territory of the former USSR. European Journal of Epidemiology 9, 86–91. Kuiper, H., de Jong, B.M., Nauta, A.P., Houweling, H., Wiessing, L.G., Moll van Charante, A.W. and Spanjaard, L. (1991) Lyme borreliosis in Dutch forestry workers. Journal of Infection 23, 279–286. Lane, R.S., Piesman, J. and Burgdorfer, W. (1991) Lyme borreliosis: relation of its causative agent to its vectors and host in North America and Europe. Annual Review of Entomology 36, 587–609. Lastavica, C.C., Wilson, M.L., Berardi, V.P., Spielman, A. and Deblinger, R.D. (1989) Rapid emergence of a focal epidemic of Lyme disease in coastal Massachusetts. New England Journal of Medicine 320, 133–137.

275

Ley, C., Le, C., Olshen, E.M. and Reingold, A.L. (1994) The use of serologic tests for Lyme disease in a prepaid health plan in California. Journal of the American Medical Association 271, 460–463. Ley, C., Olshen, E.M. and Reingold, A.L. (1995) Case–control study of risk factors for incident Lyme disease in California. American Journal of Epidemiology 142, S39–S47. Li, M., Masuzawa, T., Takada, N., Ishiguro, F., Fujita, H., Iwaki, A., Wang, H., Wang, J., Kawabata, M. and Yanagihara, Y. (1998) Lyme disease Borrelia species in northeastern China resemble those isolated from far eastern Russia and Japan. Applied and Environmental Microbiology 64, 2705–2709. Lightfoot, R.W. Jr, Luft, B.J., Rahn, D.W., Steere, A.C., Sigal, L.H., Zoschke, D.C., Gardner, P., Britton, M.C. and Kaufman, R.L. (1993) Empiric parenteral antibiotic treatment of patients with fibromyalgia and fatigue and a positive result for Lyme disease: a costeffectiveness analysis. Annals of Internal Medicine 119, 503–509. Lindenmayer, J.M., Marshall, D. and Onderdonk, A.B. (1991) Dogs as sentinels for Lyme disease in Massachusetts. American Journal of Public Health 81, 1448–1454. Logigian, E.L., Kaplan, R.F. and Steere, A.C. (1990) Chronic neurologic manifestations of Lyme disease. New England Journal of Medicine 323, 1438–1444. MacDonald, A.B., Benach, J.L. and Burgdorfer, W. (1987) Stillbirth following maternal Lyme disease. New York State Journal of Medicine 87, 615–616. Magid, D., Schwartz, B., Craft, J. and Schwartz, J.S. (1992) Prevention of Lyme disease after tick bites: a costeffectiveness analysis. New England Journal of Medicine 324, 534–541. Magnarelli, L.A. and Anderson, J.F. (1988) Ticks and biting insects infected with the etiologic agent of Lyme disease,

276

D.T. Dennis and E.B. Hayes

Borrelia burgdorferi. Journal of Clinical Microbiology 26, 1482–1486. Maraspin, V., Lotric-Furlan, S., Cimperman, J., Ruzic-Sabljic, E. and Strle, F. (1999) Erythema migrans in the immunocompromised host. Wiener Klinische Wochenschrift 111, 923–932. Marcus, L.C., Steere, A.C., Duray, P.H., Anderson, A.E. and Mahoney, E.B. (1985) Fatal pancarditis in a patient with coexistent Lyme disease and babesiosis: demonstration of spirochetes in the myocardium. Annals of Internal Medicine 103, 374–376. Marshall, W.F. III, Telford, S.R. III, Rys, P.N., Rutledge, B.J., Mathiesen, D., Malawista, S.E., Spielman, A. and Persing, D.H. (1994) Detection of Borrelia burgdorferi DNA in museum specimens of Peromyscus leucopus. Journal of Infectious Diseases 170, 1027–1032. Matuschka, F.R., Endepols, S., Richter, D. and Spielman, A. (1997) Competence of urban rats as reservoir hosts for Lyme disease spirochetes. Journal of Medical Entomology 34, 489–493. Maupin, G.O., Fish, D., Zultowsky, J., Campos, E.G. and Piesman, J. (1991) Landscape ecology of Lyme disease in a residential area of Westchester county, New York. American Journal of Epidemiology 133, 1105–1113. Meek, J.I., Roberts, C.L., Smith, E.V. Jr and Cartter, M.L. (1996) Underreporting of Lyme disease by Connecticut physicians, 1992. Journal of Public Health Management Practice 2, 61–65. Meltzer, M.J., Dennis, D.T. and Orloski, K.A. (1999) The cost effectiveness of vaccinating against Lyme disease. Emerging Infectious Diseases 5, 321–328. de Mik, E.L., van Pelt, W., Docters-van Leeuwen, B., van der Veen, A., Schellekens, J.F.P. and Borgdorff, M.W. (1997) The geographical distribution of tick bites and erythema migrans in general practice in the Netherlands. International Journal of Epidemiology 26, 451–457.

Nadelman, R.B. and Wormser, G.P. (1995) Erythema migrans and early Lyme disease. American Journal of Medicine 98 (suppl. 4A), 15S–24S. Nadelman, R.B. and Wormser, G.P. (1998) Lyme borreliosis. Lancet 352, 557–565. Nadelman, R.B., Nowakowski, J., Fish, D., Falco, R.C., Freeman, K., McKenna, D., Welch, P., Marcus, R., AgueroRosenfeld, M.E., Dennis, D.T. and Wormser, G.P. (2001) Prophylaxis with single-dose doxycycline for the prevention of Lyme disease after an Ixodes scapularis tick bite. New England Journal of Medicine 345, 79–84. Nakama, H., Muramatsu, K., Uchikawa, K. and Yamagishi, T. (1994) Possibility of Lyme disease as an occupational disease – seroepidemiological study of regional residents and forestry workers. Asia-Pacific Journal of Public Health 7, 214–217. Nakao, M., Uchikawa, K. and Dewa, H. (1996) Distribution of Borrelia species associated with Lyme disease in the subalpine forests of Nagano prefecture, Japan. Microbiology and Immunology 40, 307–311. Nichol, G., Dennis, D.T., Steere, A.C., Lightfoot, R., Wells, G., Shea, B. and Tugwell, P. (1998) Test-treatment strategies for patients suspected of having Lyme disease: a cost-effectiveness analysis. Annals of Internal Medicine 128, 37–48. O’Connell, S., Granstrom, M., Gray, J.S. and Stanek, G. (1998) Epidemiology of European Lyme borreliosis. Zentralblatt für Bakteriologie 287, 229–240. Olsen, B., Duffy, D.C., Jaenson, T.G.T., Gylfe, A., Bonnedahl, J. and Bergstrom, S. (1995) Transhemispheric exchange of Lyme disease spirochetes by seabirds. Journal of Clinical Microbiology 33, 3270–3274. Orloski, K.A., Campbell, G.L., Genese, C.A., Beckley, J.W., Schriefer, M.E., Spitalny, K.C. and Dennis, D.T. (1998) Emergence of Lyme disease in

Epidemiology of Lyme Borreliosis

Hunterdon County, New Jersey, 1993: a case-control study of risk factors and evaluation of reporting patterns. American Journal of Epidemiology 147, 391–397. Orloski, K.A., Hayes, E.B., Campbell, G.L. and Dennis, D.T. (2000) Surveillance for Lyme disease – United States, 1992–1998. Morbidity and Mortality Weekly Report 49, 1–11. Persing, D.H., Telford, S.R. III, Rys, P.N., Dodge, D.E., White, T.J., Malawista, S.E. and Spielman, A. (1990) Detection of Borrelia burgdorferi DNA in museum specimens of Ixodes dammini ticks. Science 249, 1420–1423. Petersen, L.R., Sweeney, A.H., Checko, P.J., Magnarelli, L.A., Mshar, P.A., Gunn, R.A. and Hadler, J.L. (1989) Epidemiological and clinical features of 1,149 persons with Lyme disease identified by laboratory-based surveillance in Connecticut. Yale Journal of Biology and Medicine 62, 253–262. Pfister, H.-W., Preac-Mursic, V., Wilske, B., Schielke, E., Sorgel, F. and Einhaupl, K.M. (1991) Randomized comparison of ceftriaxone and cefotaxime in Lyme neuroborreliosis. Journal of Infectious Diseases 163, 311–318. Picken, R.N., Cheng, Y., Strle, F. and Picken, M.M. (1996) Patient isolates of Borrelia burgdorferi sensu lato with genotypic and phenotypic similarities to strain 25015. Journal of Infectious Diseases 174, 1112–1115. Postic, D., Korenberg, E., Gorelova, N., Kovalevski, Y.V., Bellenger, E. and Baranton, G. (1997) Borrelia burgdorferi sensu lato in Russia and neighbouring countries: high incidence of mixed isolates. Research in Microbiology 148, 691–702. Rath, P.M., Ibershoff, B., Mohnhaupt, A., Albig, J., Eljaschewitsch, B., Jurgens, D., Horbach, I., and Fehrenbach, F.J. (1996) Seroprevalence of Lyme borreliosis in forestry workers from Brandenburg, Germany. European

277

Journal of Clinical Microbiology and Infectious Diseases 15, 372–377. Reimer, B., Marschang, A., Fingerle, V., Wilske, B., van Sonnenburg, F. and van Hoecke, C. (1999) Epidemiology of Lyme borreliosis in south-eastern Bavaria (Germany). Zentralblatt für Bakterologie 289, 653–654. Richter, D., Spielman, A., Komar, N., and Matuschka, F.R. (2000) Competence of American robins as reservoir hosts for Lyme disease spirochetes. Emerging Infectious Disease 6, 133–138. Robertson, J.N., Gray, J.S. and Stewart, P. (2000a) Tick bite and Lyme borreliosis risk at a recreational site in England. European Journal of Epidemiology 16, 647–652. Robertson, J., Guy, E., Andrews, N., Wilske, B., Anda, P., Granstrom, M., Hauser, U., Moosmann, Y., Sambri, V., Schellekens, J., Stanek, G. and Gray, J. (2000b) A European multicenter study of immunoblotting in serodiagnosis of Lyme borreliosis. Journal of Clinical Microbiology 38, 2097–2102. Saint Girons, I., Gerns, L., Gray, J.S., Guy, E.C., Korenberg, E., Nuttall, P.A., Rijpkema, S.G.T., Schonberg, A., Stanek, G. and Postic, D. (1998) Identification of Borrelia burgdorferi sensu lato species in Europe. Zentralblatt für Bakteriologie 287, 190–195. Schlesinger, P.A., Duray, P.H., Burke, P.A., Steere, A.C. and Stillman, T. (1985) Maternal–fetal transmission of the Lyme disease spirochete, Borrelia burgdorferi. Annals of Internal Medicine 103, 67–68. Schmid, G.P. (1985) The global distribution of Lyme disease. Review of Infectious Diseases 7, 41–50. Schneider, B.S., Zeidner, N.S., Burkot, T.R., Maupin, G.O. and Piesman, J. (2000) Borrelia isolates in northern Colorado identified as Borrelia bissetti. Journal of Clinical Microbiology 38, 3103–3105. Schwartz, B.S. and Goldstein, M.D. (1990)

278

D.T. Dennis and E.B. Hayes

Lyme disease in outdoor workers: risk factors, preventive measures, and tick removal methods. American Journal of Epidemiology 131, 877–884. Scoles, G.A., Papero, M., Beati, L. and Fish, D. (2001) A relapsing fever group spirochete transmitted by Ixodes scapularis ticks. Vector Borne and Zoonotic Diseases 1, 19–32. Seltzer, E.G. and Shapiro, E.D. (1996) Misdiagnosis of Lyme disease: when not to order serologic tests. Pediatric Infectious Diseases Journal 15, 762–763. Shadick, N.A., Liang, M.H., Phillips, C.B., Fossel, K. and Kuntz, K.M. (2001) The cost-effectiveness of vaccination against Lyme disease. Archives of Internal Medicine 161, 554–561. Sigal, L.H. (1994) Persisting complaints attributed to chronic Lyme disease: possible mechanisms and implications for management. American Journal of Medicine 96, 365–374. Smith, P.F., Benach, J.L., White, D.J., Stroup, D.F. and Morse, D.L. (1988) Occupational risk of Lyme disease in endemic areas of New York State. Annals of the New York Academy of Sciences 539, 289–301. Smith, R.P., Rand, P.W., Lacombe, E.H., Telford, S.R., Rich, S.M., Piesman, J. and Spielman, A. (1993) Norway rats as reservoir hosts for Lyme disease spirochetes on Monhegan island, Maine. Journal of Infectious Diseases 168, 687–691. Snydman, D.R., Schenkein, D.P., Berardi, V.P., Lastavica, C.C. and Pariser, K.M. (1986) Borrelia burgdorferi in joint fluid in chronic Lyme arthritis. Annals of Internal Medicine 104, 798–800. Sood, S.K., Salzman, M.B., Johnson, B.J.B., Happ, C.M., Fieg, K., Carmody, L., Rubin, L.G., Hilton, E. and Piesman, J. (1997) Duration of tick attachment as a predictor of the risk of Lyme disease in an area in which Lyme disease is endemic. Journal Infectious Diseases 175, 996–999. Spielman, A. (1994) The emergence of

Lyme disease and human babesiosis in a changing environment. Annals of the New York Academy of Sciences 740, 146–156. Spielman, A., Wilson, M.L., Levine, J.F. and Piesman, J. (1985) Ecology of Ixodes dammini-borne human babesiosis and Lyme disease. Annual Review of Entomology 30, 439–460. Stafford, K.C., Cartter, M.L., Magnarelli, L.A., Ertel, S.H. and Mshar, P.A. (1998) Temporal correlations between tick abundance and prevalence of ticks infected with Borrelia burgdorferi and increasing incidence of Lyme disease. Journal of Clinical Microbiology 36, 1240–1244. Stanek, G., Klein, J., Bittner, R. and Glogar, D. (1990) Isolation of Borrelia burgdorferi from the myocardium of a patient with longstanding cardiomyopathy. New England Journal of Medicine 322, 249–252. Stanek, G., O’Connell, S., Cimmino, M., Aberer, E., Kristoferitsch, W., Granstrom, M., Guy, E. and Gray, J. (1996) European Union Concerted Action on risk assessment in Lyme borreliosis: clinical case definitions for Lyme borreliosis. Weiner Klinische Wochenschrift 108, 741–747. Steere, A.C. (1989) Medical progress: Lyme disease. New England Journal of Medicine 321, 586–596. Steere, A.C. (2001) Lyme disease. New England Journal of Medicine 345, 115–125. Steere, A.C., Malawista, S.E., Snydman, D.R., Shope, R.E., Andiman, W.A., Ross, M.R. and Steele, F.M. (1977) Lyme arthritis: an epidemic of oligoarticular arthritis in children and adults in three Connecticut communities. Arthritis and Rheumatism 20, 7–17. Steere, A.C., Broderick, T.F. and Malawista, S.E. (1978) Erythema chronicum migrans and Lyme arthritis: epidemiological evidence for a tick vector. American Journal Epidemiology 108, 685–695.

Epidemiology of Lyme Borreliosis

Steere, A.C., Taylor, E., Wilson, M.L., Levine, J.F. and Spielman, A. (1986) Longitudinal assessment of the clinical and epidemiological features of Lyme disease in a defined population. Journal of Infectious Diseases 154, 295–300. Steere, A.C., Taylor, E., McHugh, G.L. and Logigian, E.L. (1993) The overdiagnosis of Lyme disease. Journal of the American Medical Association 269, 1812–1816. Steere, A.C., Levin, R.E., Molloy, P.J., Kalish, R.A., Abraham, J.H., Liu, N.Y. and Schmid, C.H. (1994) Treatment of Lyme arthritis. Arthritis and Rheumatism 37, 878–888. Steere, A.C., Sikand, V.K., Meurice, F., Parenti, D.L., Fikrig, E., Schoen, R.T., Nowakowski, J., Schmid, C.H., Laukamp, S., Buscarino, C., Krause, D.S. and the Lyme Disease Vaccine Study Group (1998) Vaccination against Lyme disease with recombinant Borrelia burgdorferi outer-surface lipoprotein A with adjuvant. New England Journal of Medicine 339, 209–215. Stobierski, M.G., Hall, W.N., RobinsonDunn, B., Stiefel, H., Shiflet, S. and Carlson, V. (1994) Isolation of Borrelia burgdorferi from two patients in Michigan. Clinical Infectious Diseases 19, 944–946. Strickland, G.T., Karp, A.C., Mathews, A. and Pena, C.A. (1997) Utilization and cost of serological tests for Lyme disease in Maryland. Journal of Infectious Diseases 176, 819–821. Strle, F. (1999) Lyme borreliosis in Slovenia. Zentralblatt für Bakteriologie 289, 643–652. Strle, F., Cheng, Y., Cimperman, J., Maraspin, V., Lotric-Furlan, S., Nelson, J.A., Picken, M.M., Ruzic-Sabljic, E. and Picken, R.N. (1995) Persistence of Borrelia burgdorferi sensu lato in resolved erythema migrans lesions. Clinical Infectious Diseases 21, 380–389. Strobino, B.A., Williams, C.L., Abid, S., Chalson, R. and Spierling, P. (1993)

279

Lyme disease and pregnancy outcome: a prospective study of two thousand prenatal patients. American Journal of Obstetrics and Gynecology 169, 367–374. Strobino, B.A., Abid, S. and Gewitz, M. (1999) Maternal Lyme disease and congenital heart disease: a case–control study in an endemic area. American Journal of Obstetrics and Gynecology 180, 711–716. Talaska, T. (1998) BorrelioseEpidemiologie am Biespiel des Bundeslandes Brandenburg. Für Die Praxis: Lyme-Borreliose 40–41. Trevejo, R.T., Krause, P.J., Schriefer, M.E. and Dennis, D.T. (2001) Evaluation of a two-test serodiagnostic method for community assessment of Lyme disease in an endemic area. American Journal of Tropical Medicine and Hygiene 65, 563–566. Tugwell, P., Dennis, D.T., Weinstein, A., Wells, G., Shea, B., Nichol, G., Hayward, R., Lightfoot, R., Baker, P. and Steere, A.C. (1997) Laboratory evaluation in the diagnosis of Lyme disease. Annals of Internal Medicine 127, 1109–1123. des Vignes, F., Piesman, J., Heffernan, R., Schulze, T.L., Stafford, K.C. and Fish, D. (2001) Effect of tick removal on transmission of Borrelia burgdorferi and Ehrlichia phagocytophila by Ixodes scapularis nymphs. Journal of Infectious Diseases 183, 773–778. Wanchun, T., Zhikun, Z., Moldenhauer, S., Yixiu, G., Qiuli, G., Qiuli, Y., Lina, W. and Mei, C. (1998) Detection of Borrelia burgdorferi from ticks (Acari) in Hebei Province, China. Journal of Medical Entomology 35, 95–98. Wang, G., van Dam, A.P., Schwartz, I. and Dankert, J. (1999) Molecular typing of Borrelia burgdorferi sensu lato: taxonomic, epidemiological, and clinical implications. Clinical Microbiology Reviews 12, 633–653. Warshafsky, S., Nowakowski, J., Nadelman, R.B., Kamer, R.S., Peterson, S.J. and

280

D.T. Dennis and E.B. Hayes

Wormser, G.P. (1996) Efficacy of antibiotic prophylaxis for prevention of Lyme disease. Journal of General Internal Medicine 11, 329–333. Weber, K. (2001) Aspects of Lyme borreliosis in Europe. European Journal of Clinical Microbiology and Infectious Diseases 20, 6–13. White, D.J., Chang, H., Benach, J.L., Bosler, E.M., Meldrum, S.C., Means, R.G., Debbie, J.G., Birkhead, G.S. and Morse, D.L. (1991) The geographic spread and temporal increase of the Lyme disease epidemic. Journal of the American Medical Association 266, 1230–1236.

Wilson, M.L. (1998) Distribution and abundance of Ixodes scapularis (Acari: Ixodidae) in North America: ecological processes and spatial analysis. Journal of Medical Entomology 35, 446–457. Wormser, G.P., Bittker, S., Cooper, D., Nowakowski, J., Nadelman, R.B. and Pavia, C. (2000) Comparison of the yields of blood cultures using serum or plasma from patients with early Lyme disease. Journal of Clinical Microbiology 38, 1648–1650. Yanagihara, Y. and Masuzawa, T. (1997) Lyme disease (Lyme borreliosis). FEMS Immunology and Medical Microbiology 18, 249–261.

Vaccination Against Lyme Borreliosis

11

Edward B. Hayes and Martin E. Schriefer Division of Vector-borne Infectious Diseases, Centers for Disease Control and Prevention, Public Health Service, US Department of Health and Human Services, PO Box 2087, Fort Collins, CO 80522, USA

Introduction Within 20 years of the identification of the tick vectors of Lyme borreliosis (LB) and isolation of the North American causative agent, Borrelia burgdorferi sensu stricto (s.s.), an efficacious vaccine to prevent LB in humans was commercially available. Vaccine development began with the demonstration that both passive and active immunization of hamsters protected these animals against intraperitoneal challenge with B. burgdorferi s.s. By 1990, a chemically inactivated wholecell vaccine was being administered to dogs, although the efficacy of the vaccine had not been convincingly demonstrated. Whole-cell vaccines were not developed for humans because of concerns that such preparations might induce deleterious autoimmune reactions. Development of vaccines for humans focused on immunization with outer surface protein A (OspA) following the discovery that this protein was relatively homogenous among B. burgdorferi s.s. strains isolated from North America, and that immunization with OspA was highly protective against B. burgdorferi s.s. infection in mice. Two pharmaceutical companies developed recombinant OspA vaccines in the early 1990s and conducted independent phase III trials with over 10,000 participants in each trial. One of these vaccines, LYMErixTM, is currently licensed for use and until recently was commercially available in the USA*. The efficacy of the vaccine in protecting against LB in Europe or Asia is unknown. *Note: The manufacturer of LYMErixTM, the only commercially available vaccine against Lyme disease, discontinued sales of the vaccine in February 2002 because of low commercial demand for the vaccine. The authors are not aware of any current plans to resume manufacture or distribution of the vaccine. The recommendations for use of the vaccine at the time of discontinuation of sales were unchanged from those described in this chapter. © 2002 CAB International. Lyme Borreliosis: Biology, Epidemiology and Control (eds J. Gray, O. Kahl, R.S. Lane and G. Stanek)

281

282

E.B. Hayes and M.E. Schriefer

The impetus for development of a vaccine against LB for humans arose from the increasing numbers of LB cases reported to health authorities in the USA, and the increased recognition of LB as an important vector-borne disease in Europe and temperate Asia. In addition, although the disease is rarely if ever fatal and responds well to treatment if diagnosed in its early stages, failure to receive early treatment can sometimes result in disabling sequelae, which may be refractory to antibiotic therapy. Efforts to prevent LB by educating people to wear protective clothing and repellents, and to check for and remove attached ticks, while theoretically efficacious, have not been effective in reducing the incidence of disease in endemic areas. Proposals to reduce vector tick populations in endemic areas by applying acaricides to tick habitats have generated concern about potentially toxic effects of the acaricides on the environment. The elimination of deer can reduce tick densities in isolated endemic foci, but this intervention has not been adopted on a wide scale. Thus, in the early 1990s, efforts were focused on developing an effective vaccine to prevent LB.

Epidemiology of Lyme Borreliosis in Relation to Vaccination Reliable estimates of LB case numbers are difficult to obtain but approximately 17,000 new cases are reported annually in the USA and it has been estimated that in ten European countries nearly 60,000 cases may occur each year (see Dennis and Hayes, Chapter 10). In the USA, a geographical analysis of LB risk based on reported cases of LB, distribution of vector ticks and impact of host distribution on the prevalence of infection in ticks identified 115 (4%) counties in the USA as being high risk for LB, 146 (5%) as moderate risk, 1143 (36%) as low risk and 1736 (55%) as minimal or no risk (Centers for Disease Control and Prevention, 1999). The high-risk counties are all within the northeast, mid-Atlantic and northcentral regions of the country. This focal distribution of LB risk within the USA provided the basis for public health recommendations for vaccine use that target the vaccine to people who are likely to be exposed to tick habitat in high or moderate risk areas (Centers for Disease Control and Prevention, 1999). LB risk in Europe appears to increase from west to east, paralleling the prevalence of infected ticks, with the highest reported incidence in Austria and Slovenia, but foci of high risk have also been described in western Europe (O’Connell et al., 1998). The incidence of LB in the USA is highest in children aged 5–9 years and adults aged 45–54 years (Orloski et al., 2000). A study in Sweden found the highest incidence in children aged 5–9 years and adults aged 60–74 years (Berglund et al., 1995). The relatively high risk of LB in children, as well as the potential for children to fail to notice or report to their parents an erythema migrans (EM) rash, have prompted efforts to develop a vaccine that will be safe and efficacious in children as well as adults. Males and females appear to be at approximately equal risk in the USA, but in Europe both equal risk by gender and higher risk in males have been reported (Berglund et al., 1995; O’Connell et al., 1998). LB

Vaccination Against Lyme Borreliosis

283

risk may be higher than average in some occupational and recreational groups, which have more frequent contact with infected ticks (Smith et al., 1988; O’Connell et al., 1998). Another spirochaete, B. lonestari, has been associated with the occurrence of a rash, strongly resembling the characteristic EM of LB, in humans in southern and central areas of the USA that are not endemic for LB (see Dennis and Hayes, Chapter 10). An ospA homologue has not been identified in B. lonestari, despite attempts to amplify ospA from ticks (Amblyomma americanum) and from EM biopsies (Felz et al., 1999; James et al., 2001), which suggests that vaccination with recombinant OspA (rOspA) is unlikely to provide protection against this spirochaete.

Development of Recombinant OspA Vaccines Despite the early development of whole-cell vaccines for dogs, there was concern that whole cell vaccines in humans would be highly reactogenic and might elicit immune responses that would cross-react with human antigens and provoke inflammatory tissue damage (Wormser, 1995). Research towards developing vaccines in humans therefore focused on subunit vaccines. Several proteins expressed by B. burgdorferi s.s. were considered as candidates for vaccine development, primarily OspA, OspB and OspC. These outer surface proteins were shown to be protective in animal models (Probert and LeFebvre, 1994). There was some indication that both OspA and OspB conferred protection against infection and overt disease (Fikrig et al., 1990; Telford and Fikrig, 1995). This was seen as a desirable property for a candidate vaccine, because of concerns that subclinical infection might result in late-stage sequelae such as those seen in untreated naturally acquired LB. Because of the relative homogeneity of OspA in B. burgdorferi s.s. isolates (Fikrig et al., 1990), efforts to develop a human subunit vaccine focused on this protein. Despite the finding that immunization with OspA could provide protection against infection in mice, protective immunity was achieved only against homologous B. burgdorferi s.s. strains. Two important protective and non-overlapping B cell epitopes defined by the monoclonal antibodies LA-2 and 336 are located at the C-terminal end of OspA (Kalish et al., 1995; Wormser, 1995; Ding et al., 2000). Protective immunity is more highly correlated with levels of LA-2-specific antibody than with total levels of OspA antibody. LA-2 recognizes three exposed loops on the surface of B. burgdorferi s.s. OspA, but does not bind to this region of Borrelia garinii or Borrelia afzelii OspA. Amino acid variation within these loops may account for the failure of B. burgdorferi s.s. OspA to elicit protection against heterologous Borrelia spp. and strains. None the less, variation within this looped epitope of B. garinii and B. afzelii strains is limited, suggesting that a multivalent OspA vaccine that provides broader protection against all three borrelial genospecies may be feasible (Telford and Fikrig, 1995; Ding et al., 2000). An N-terminal lipid moiety enhances immunogenicity of OspA (Erdile et al., 1993; van Hoecke et al., 1996).

284

E.B. Hayes and M.E. Schriefer

While in the vector tick midgut, B. burgdorferi sensu lato (s.l.) expresses OspA, but after migration from the tick’s gut to the salivary glands and on entering the mammalian host, the expression of OspA is suppressed and most spirochaetes preferentially express OspC (de Silva et al., 1996). Vaccination with OspA elicits antibody that destroys or inactivates the spirochaetes while they are still inside the tick’s gut. This protective mechanism is unique among human vaccines. Presumably, any spirochaetes that enter host tissue while still expressing OspA can be neutralized through a more traditional antibody–antigen interaction within the vaccinated host’s tissues (Telford and Fikrig, 1995). The actual immune mechanism by which antibody to OspA provides protection is not known, but may involve enhanced phagocytosis through Fc receptor binding (Montgomery et al., 1994), immobilizing, killing or inhibiting the growth of spirochaetes (Pavia et al., 1991; Keller et al., 1994; Wormser, 1995), or blocking of spirochaete attachment to cell surfaces (Wormser, 1995). In the early 1990s, Kalish et al. (1993) determined that individuals with treatment-resistant Lyme arthritis following naturally acquired infection were more likely to have detectable antibody to OspA and a higher prevalence of certain HLA-DR4 haplotypes than those patients whose Lyme arthritis responded to conventional antibiotic therapy. An autoimmune mechanism was proposed as the cause for the persistent joint inflammation and these findings generated concerns that OspA vaccines could induce autoimmune reactions. Further research indicated that certain epitopes of OspA were structurally similar to epitopes on human leukocyte function-associated antigen-1 and that this similarity might provide the basis for an autoimmune-mediated release of inflammatory cytokines in joint tissue (Gross et al., 1998). Whether vaccination with OspA could provoke or exacerbate such a reaction remains unknown, but the results of several clinical trials in humans, described below, suggest that immunization alone does not cause arthritis.

Clinical Trials in Humans The immunogenicity and safety of the vaccine in humans was first evaluated in 24 healthy volunteers, with no known exposure to LB, aged 18–65 in the nonendemic state of New Mexico (Keller et al., 1994). Twelve participants received 10 µg recombinant OspA without adjuvant and 12 received 10 µg rOspA with 0.5 mg aluminium adjuvant. An additional 12 participants received a placebo. The most common adverse event was tenderness or pain at the injection site, which occurred in 22 (46%) persons after the first dose of rOspA, in 20 (42%) persons after the second dose and in eight (67%) of 12 persons who were given a third dose of vaccine without adjuvant. Local reactions occurred more frequently in vaccinated persons than among placebo recipients, but the study lacked sufficient power to detect a significant difference. No serious adverse events were reported. Systemic reactions such as headache within 72 h, fatigue within 6 h and fever within 24 h were reported after each injection by fewer than 20%

Vaccination Against Lyme Borreliosis

285

of vaccine and placebo recipients. Seven vaccine recipients and one placebo recipient reported joint pain. In all cases the joint pain resolved without treatment and the investigators believed it to be causally unrelated to vaccination. Joint pain was more frequently reported by vaccine and placebo recipients who were HLA-DR4-positive, but this difference was not statistically significant. None of the participants developed a clinical picture consistent with Lyme arthritis. Both vaccine formulations produced high titres of antibodies that inhibited B. burgdorferi s.s. growth in vitro. The geometric mean titres of antibody elicited by the two formulations were similar and were significantly higher than in the group given placebo. The safety and immunogenicity of three different dosage strengths of rOspA vaccine with adjuvant was evaluated in an uncontrolled clinical trial in adults with previous LB (Schoen et al., 1995). Five participants received 5 µg of the vaccine, five received 10 µg, and 20 received 30 µg. All but two of the participants received three doses of vaccine at 0, 1 and 2 months. No serious adverse events and no sustained systemic adverse effects were reported. Local reactions at the injection site were reported frequently, but participants rated 96% of these reactions as mild. The most common reaction was soreness, reported by 40–85% of vaccine recipients after any given dose. Systemic reactions after injection included fever and rash (each reported by no more than 20% of participants after any given dose), headache and fatigue (each reported by up to 40% of participants after any given dose) and arthralgia (reported by no more than 33% of participants after any given dose). Eighty-one per cent of these reactions were classified as mild. All adverse events reported in the study were of short duration. All dosage strengths of the vaccine resulted in increased geometric mean titres of OspA antibody, but those receiving the 30 µg dose developed higher titres than those receiving the lower dose formulations. The occurrence of adverse reactions was not related to antibody titre. The safety and immunogenicity of three different formulations of rOspA vaccine was assessed in 300 seronegative volunteers in a randomized trial without a placebo group (van Hoecke et al., 1996). The three vaccine formulations were non-structural protein (NS1)–OspA, NS1–OspA with monophosphoryl lipid A (MPL) and lipoprotein OspA, all with 10 µg of rOsp A adsorbed to aluminium hydroxide. Vaccines were administered on days 0, 28 and 56. In the first component of the study, each formulation was given to 20 different participants. At least one symptom, mostly mild local reactions, was reported by 35–47% of vaccine recipients in each group. All symptoms resolved within 4 days. Fever, headache and malaise were reported by fewer than 12% of participants and none of these general symptoms was felt to be related to the vaccine. One person developed pneumonia that was unrelated to the vaccine. No other severe adverse events were reported. In the second component of the study, each formulation was administered to 80 different participants. Pain at the injection site was the most frequent local reaction, reported after 24–49% of injections. General symptoms such as headache, fever or malaise were reported by 10% or fewer participants after any given dose, and were reported as either mild or

286

E.B. Hayes and M.E. Schriefer

moderate. Two participants reported mild arthralgia that resolved within 3 days of receiving their first vaccine dose. No severe adverse events were reported. The lipoprotein formulation produced a higher and more persistent antibody response than the other two formulations. Following the generally favourable results of these phase I and phase II trials, two vaccine manufacturers sponsored large phase III trials of OspA vaccine to assess safety, immunogenicity and efficacy in preventing LB. Pasteur Merieux Connaught sponsored a trial of rOspA lipoprotein vaccine without adjuvant (Sigal et al., 1998). The lipoprotein was derived from a B31 strain of B. burgdorferi s.s. that was expressed by Escherichia coli. The vaccine was given to 5156 people 18 years of age or older and an additional 5149 participants received placebo. Sigal et al. (1998) reported a vaccine efficacy of 68% (95% CI 36%–85%) after two doses of vaccine in the first year. However, the criteria for confirming the diagnosis of LB in study participants were unclear and only a subset of suspected cases were reviewed in depth to confirm diagnosis. Data on efficacy after a third dose were reported but administration of the third dose was not designed into the original protocol and was only given to a subset of participants selected through unspecified criteria. There were no serious adverse events associated with the vaccine in this trial. Vaccine recipients reported pain and inflammation at the injection site significantly more frequently than did placebo recipients. SmithKline Beecham sponsored a randomized clinical trial to assess the safety, immunogenicity and efficacy of rOspA lipoprotein vaccine with aluminium hydroxide adjuvant (Steere et al., 1998). The ospA gene of B. burgdorferi s.s. strain ZS7 was expressed in transformed E. coli. The vaccine was administered to 5469 study participants aged 15–70 years, and 5467 participants of the same age range received a placebo consisting of adjuvant in buffered saline. Doses of vaccine and placebo were administered at 0, 1 and 12 months and timed to provided high antibody levels just before the late spring increase in LB transmission. The following criteria were used to diagnose LB in both study groups: the presence of investigator-observed EM or objective neurologic, musculoskeletal or cardiovascular manifestations of LB, with laboratory confirmation of infection by culture, PCR, or seroconversion documented by Western blot. The estimated vaccine efficacy in reducing laboratory-confirmed LB was 49% (95% CI 15%–69%, NNT 256, where NNT = the number needed to treat, or in this case, the number of persons that would have to be vaccinated to prevent one case of LB) after two doses in the first year and 76% (95% CI 58%–86%, NNT 109) after a third dose administered 1 year later. The efficacy in preventing asymptomatic infection was 83% (95% CI 32%–97%, NNT 500) after two doses, and 100% (95% CI 26%–100%, NNT 370) after three doses. The development of OspA antibody was evaluated in a subset of participants from a single study site (SmithKline Beecham Pharmaceuticals, 1998, LYMErixTM product label). Of approximately 460 participants who received vaccine at the site, serologic results were reported for 267 participants. The criteria for selection of these participants were not specified. The geometric mean titre (GMT) was 1227 ELISA units ml–1 at month 2 (n = 264, 1 month after the second dose), 116 ELISA units

Vaccination Against Lyme Borreliosis

287

ml–1 at month 12 before the third dose (n = 241), 6006 ELISA units ml–1 at month 13 (n = 267, 1 month after the third dose) and 1991 ELISA units ml–1 at month 20 (n = 267). ELISA units ml–1 were defined based on an evaluation of pooled sera from people who had been vaccinated with OspA (van Hoecke et al., 1996). A subsequent analysis by the manufacturer concluded that a titre greater than 1400 ELISA units ml–1 was associated with protection from disease (van Hoecke et al., 1999). Data on adverse events in this study were collected in two separate ways. At one site 938 participants were asked to complete diary cards recording adverse events for the first 3 days after each dose. The remaining 9998 participants at other study sites were simply asked to report any adverse events to the investigators. Among the 9998, the most frequently described adverse event that was determined to be related or possibly related to the vaccine was soreness at the injection site, which was reported by 24.1% of vaccine recipients compared with 7.6% of placebo recipients (P < 0.001, NNH 6, where NNH = number needed to harm, or in this case, the number of persons that would be vaccinated for every single occurrence of the specified adverse event). Systemic events were reported by 19.4% of vaccine recipients compared with 15.1% of placebo recipients (P < 0.001, NNH 23), but no single systemic event was reported by more than 3.9% of vaccine recipients. The individual systemic events that were significantly associated with the vaccine were myalgias (3.2% of vaccine recipients versus 1.8% of placebo recipients, NNH 71), influenza-like illness (2.0% versus 1.1%, NNH 111), fever (2.0% versus 0.8%, NNH 83) and chills (1.8% versus 0.5%, NNH 77). Among the 938 participants from whom adverse event data were solicited by diary card, systemic events were reported by 63% of vaccine recipients, compared with 53% of placebo recipients (P = 0.004, NNH 10). For all 10, 936 participants, there was no significant difference between vaccine and placebo recipients in the proportions reporting adverse events 30 days or more after receiving a dose at the P = 0.05 significance level. The incidence of arthritis was not significantly different between vaccine and placebo recipients. No hypersensitivity reactions were noted in vaccine recipients. A recombinant bacille Calmette-Guérin vaccine expressing OspA lipoprotein was evaluated in a dose-escalation study in 24 adults (Edelman et al., 1999). The vaccine elicited no detectable antibody to OspA in any of the volunteers and induced skin ulcers at the higher dose strength.

Clinical trials in children The safety and immunogenicity of rOspA lipoprotein vaccine were assessed in a clinical trial that randomly assigned 250 children aged 5–15 years in the Czech Republic to receive either 15 µg or 30 µg of the vaccine (Feder et al., 1999). There was no placebo group. Each study group received three doses of vaccine on a 0, 1 and 2 month schedule. None of the 250 children had detectable antibody to OspA at the time they were enrolled in the study. One month after the

288

E.B. Hayes and M.E. Schriefer

second injection, the GMT of the group that received 15 µg was 2346 ELISA units ml–1, and the GMT of the group that received 30 µg was 4259 ELISA units ml–1. At month 3, after the third injection, the GMTs were 5967 ELISA units ml–1 in the 15 µg group and 10267 ELISA units ml–1 in the 30 µg group. Comparing these results with those of the phase III trial in adults (Steere et al., 1998) suggests that immunization with 30 µg of OspA tends to elicit higher OspA titres in children than in adults. Participants in this children’s trial were asked to record adverse events on diary cards for 3 days after each vaccination, and cards were returned for all children in the study after all doses. Soreness at the injection site was the most frequently reported adverse event, reported by 72% of those receiving their first 30 µg dose and 69% of those receiving their first 15 µg dose. Soreness was less frequently reported in both groups after subsequent doses. Redness at the injection site was reported by, at most, 43% of participants after any given dose. Headache and malaise were the most frequently reported systemic symptoms, with the highest proportion reported by those who received the first 15 µg dose (18% reported headache and 20% reported malaise). All systemic symptoms resolved within 6 days of vaccination and no serious adverse events were reported. The safety and immunogenicity of OspA vaccine were evaluated in 91 children aged 2–5 years in the Czech Republic, who were randomly assigned to receive two doses, 1 month apart, of either 15 µg or 30 µg of lipoprotein OspA with adjuvant (Beran et al., 2000). Pain at the injection site was the most frequently reported symptom, reported by 64% of children who received the 15 µg formulation and 69% of those receiving 30 µg. There was no indication that the 30 µg dose was substantially more reactogenic than the 15 µg dose, but the GMTs of antibody to OspA elicited by the 30 µg dose were significantly higher (9877 at 1 month after two 30 µg doses compared with 4366 after two 15 µg doses). In a randomized placebo-controlled trial to evaluate the safety and immunogenicity of OspA vaccine in children 4 years of age and older in the USA, 3063 children were vaccinated with 30 µg of rOspA vaccine and 1024 were given placebo (Sikand et al., 2001). Doses of vaccine and placebo were administered on a 0, 1 and 12 month schedule. Vaccine recipients reported a higher incidence of local reactions (78% with vaccine versus 55% with placebo; NNH 4) and general symptoms (30% with vaccine versus 22% with placebo; NNH 13) such as fatigue, fever, headache and arthralgia. The mean duration of these adverse events was 2–3 days. Study participants were asked about any rheumatological adverse events, and reports of arthritis were not any higher among vaccine recipients than among placebo recipients. The GMT of antibody to OspA 1 month after the third dose was 27,485 ELISA units ml–1 compared with a GMT of 8216 ELISA units ml–1 found in retested sera from adults in the adult Phase III efficacy trial. After three doses of vaccine, all of the children in the pediatric trial had an antibody level greater than or equal to 1400 ELISA units ml–1 compared with 90% of vaccine recipients from the adult trial.

Vaccination Against Lyme Borreliosis

289

Evaluation of alternative dosing schedules and booster doses A randomized trial of alternative dosing schedules involving 800 healthy adult volunteers found that the GMT of OspA antibody 1 month after a 0, 1 and 6 month dosing schedule was 7205 ELISA units ml–1, compared with 10,659 ELISA units ml 1 month after the 0, 1, 12-month dosing schedule (van Hoecke et al., 1999). There were no serious adverse events thought to be related to vaccination in either group. The frequency of local reactions and systemic symptoms after vaccination was similar between the two groups. Headache and malaise were the most frequently reported systemic symptoms, each reported by less than 12% of people in each group. Ninety-two per cent of participants who received the vaccine on a 0, 1 and 6 month dose schedule had titres greater than or equal to 1400 ELISA units ml–1 after the third dose, compared with 93% of those on the 0, 1 and 12 month schedule. Despite the substantial differences in the GMTs, the authors concluded that the two vaccination schedules were equivalent, because of the similarity in proportions with titres greater than or equal to 1400 ELISA units ml–1 and because the relative difference in GMTs was less than 100%. Another trial randomly assigned 956 volunteers aged 17–72 years to receive 30 µg of rOspA vaccine on either a 0, 1 and 12 month schedule or a 0, 1, 2 and 12 month schedule (Schoen et al., 2000). Ninety per cent of volunteers who were vaccinated by the 0, 1, 2 and 12 month schedule had antibody levels greater than or equal to 1400 ELISA units ml–1 after the third dose at 2 months. This group had higher GMTs after the 12 month dose than participants who were vaccinated by the 0, 1 and 12 month schedule. The adverse event profile of the two schedules after the 12 month dose was similar, with only one volunteer in each group experiencing a serious adverse event possibly related to vaccination (shaking chills in one person and syncope with full recovery in another). Studies of the immunogenicity of OspA vaccines have indicated that antibody levels to OspA decline with increasing time following vaccination. The trial by Schoen et al. (2000) indicated that a booster dose after initial immunization on a 0, 1 and 2 month schedule was needed for protection in the second year, and that the booster dose was well tolerated and provided over 99% of recipients with presumably protective antibody levels. An unpublished evaluation of the need for booster doses after a 0, 1 and 12 month initial immunization schedule suggested that booster doses were needed 1 year after the third dose (D. Parenti et al., unpublished data). The optimal schedule for booster doses in persons vaccinated on accelerated (0, 1 and 2 month, or 0, 1 and 6 month) schedules is uncertain, but it seems likely that such individuals would require a booster dose just before the tick transmission season that follows the year of their initial vaccination. Subsequent booster doses may need to be given annually or perhaps every other year, but data to answer this question and to evaluate the safety of such repeated boosters have not been published.

290

E.B. Hayes and M.E. Schriefer

Post-licensure safety and effectiveness evaluations In the USA, the Centers for Disease Control and Prevention (CDC) and the Food and Drug Administration (FDA) jointly monitor reports of adverse events following vaccinations through the Vaccine Adverse Events Reporting System (VAERS). This is a passive surveillance system that provides for collection of adverse events following vaccinations reported by patients, healthcare providers and vaccine manufacturers. The data in VAERS do not directly allow for determining causal relationships between reported adverse events and the vaccines administered, but periodic review of the data can show unexpected patterns of adverse events associated with a given vaccine and may identify rare adverse reactions to vaccines that were not seen in the limited scope of clinical trials. The reports to VAERS associated with the commercially licensed LB vaccine, LYMErixTM, from the date of its licensure in December 1998 through to July 2000, were reviewed by the CDC and FDA (Lathrop et al., 2002). During this time over 1,400,000 doses were distributed and 905 adverse events were reported. The review detected no unexpected pattern of adverse events compared with the clinical trials except that rare hypersensitivity reactions including urticaria and occasional cases of dyspnoea were reported. Such reactions had not been reported during use of the vaccine in the phase III clinical trial. The numbers of reports of arthritis were well below what would have been expected based on rates of arthritis reported in the phase III trial. A phase IV study to evaluate safety and effectiveness of the vaccine among members of managed care organizations in the northeastern USA is being conducted by the manufacturer. At least two case–control studies to evaluate postlicensure effectiveness of the vaccine in endemic states of the USA are in progress.

Utility and effectiveness of the vaccine The clinical trials of OspA vaccine have demonstrated its efficacy in preventing LB and indicated that it has an acceptable safety profile. For any given individual, the question of whether or not to be vaccinated depends on the individual’s risk of acquiring LB, the likelihood of obtaining appropriate treatment if LB is contracted, the cost and relative convenience of vaccination, and the individual’s level of concern about the consequences of acquiring LB compared with concern about possible adverse events from the vaccine. The public health benefit of vaccination to prevent LB is somewhat limited compared with many other vaccine-preventable diseases. Because LB is not contagious and humans are dead-end hosts of B. burgdorferi s.s., vaccination of some humans does not decrease the risk of LB for others. Furthermore, vaccination does not protect against other diseases, such as human granulocytic ehrlichiosis and babesiosis, which are transmitted by the tick vectors of B. burgdorferi s.l. If vaccination were to induce complacency towards exposure to ticks, then vaccinated persons could potentially be at increased risk of other tick-borne diseases. The efficacy of the vaccine

Vaccination Against Lyme Borreliosis

291

in protecting against LB outside North America is unknown, but would probably be lower because of the heterogeneity of the Borrelia spp. that cause LB in Europe and Asia. Despite these limitations, vaccination is currently one of the most effective measures available to prevent LB in the USA. Optimal protection against LB requires three doses of the vaccine, which, under current recommendations, are given over the course of a year, as they were administered in the phase III clinical trial (Centers for Disease Control and Prevention, 1999). The vaccine is administered as a 0.5 ml dose of 30 µg rOspA given by injection into the deltoid muscle. A second dose is administered 1 month after the first, and a third dose 1 year after the first dose. The administration of the vaccine should be timed such that the second dose in year 1 and the third dose in year 2 are given several weeks before the onset of the B. burgdorferi s.s. transmission season in endemic areas, which usually begins in April. This will ensure that the vaccine recipient’s OspA antibody levels will be highest during the months when the risk of acquiring LB is highest. The currently recommended schedule is somewhat awkward since it requires three doses of vaccine given in 2 separate years to achieve optimal protection. The safety and immunogenicity trial of the vaccine in Czech children indicated that high antibody levels could be achieved after three doses of vaccine given over 2 months (a 0, 1 and 2 month schedule). In adults, the 0, 1 and 2 month and 0, 1 and 6 month schedules appear to provide higher levels of antibody in the first vaccination year than the 0, 1 and 12 month schedule (van Hoecke et al., 1999; Schoen et al., 2000). Implementation of these accelerated schedules would enhance the clinical and public health utility of the vaccine since most vaccine recipients would not be exposed to a transmission season with suboptimal levels of protective antibody. However, the safety and protective efficacy of such accelerated schedules have not been fully evaluated by the FDA. The FDA has licensed LYMErixTM for use in individuals aged 15–70 years. The Advisory Committee on Immunization Practices (ACIP) of the CDC has published the following recommendations for the use of LB vaccine (CDC, 1999): 1. Vaccination should be considered for persons aged 15–70 years who engage in activities that result in frequent or prolonged exposure to tick-infested habitats in areas of high or moderate risk of acquiring LB. 2. Vaccination may be considered for persons aged 15–70 years who are exposed to tick-infested habitats in areas of high or moderate LB risk, but whose exposure is neither frequent nor prolonged. 3. Vaccination is not recommended for persons who have minimal or no exposure to tick-infested habitats, for persons in areas of low or no risk of LB, or for pregnant women. In addition, the ACIP states that the vaccine should not be given to persons with treatment-resistant Lyme arthritis because of the association of this condition with immune reactivity to OspA. The cumulative incidence of LB among placebo recipients during the second year of the phase III trial of LYMErixTM was 1.2%. Among a population with comparable incidence, and assuming a vaccine efficacy after three doses

292

E.B. Hayes and M.E. Schriefer

of 76%, about 110 people would need to be vaccinated to prevent one case of LB. The effectiveness of vaccination has been compared with other strategies for preventing LB through a model involving a hypothetical community (Hayes et al., 1999). Under the baseline assumptions of the model (which assumed a vaccine efficacy of 68% after the first dose and 92% after the third dose), vaccination was among the three most effective interventions in preventing LB. An evaluation of the cost-effectiveness of vaccination from a societal perspective indicated that at an assumed cost of vaccination of $100 per person per year, an effectiveness of 85%, a probability of 0.85 of correctly identifying and treating early LB, and an assumed cumulative incidence of LB of 1% per year, the net cost of vaccination was $5692 per case of LB prevented (Meltzer et al., 1999). Under the same assumptions the cost of vaccination exceeded the cost of not vaccinating unless the cumulative incidence of LB was above 1.97% per year. Variations in LB incidence and in the likelihood of early diagnosis and treatment each had a substantial impact on the relative cost-effectiveness of vaccination. In a separate analysis that assumed a vaccination cost of $150, vaccine efficacy of 63% after two doses and 87% after three doses, and an LB incidence of 1% per year, vaccination of 10,000 individuals prevented 202 cases over a 10 year period, and cost $62,300 per quality-adjusted life year (QALY) gained compared with no vaccination (Shadick et al., 2001). An accelerated 2 month dosing schedule reduced the cost of vaccination to $53,700 per QALY gained.

Development of Alternative Vaccines As noted in the above discussion, the recombinant OspA vaccine appears to be efficacious and have an acceptable safety profile but for several reasons it is not an ideal vaccine to prevent LB. First, the published efficacy is less than 80%, leaving 20% of vaccinated individuals susceptible to disease. Efficacy against infection with B. garinii or B. afzelii would probably be even lower because of the heterogeneity of outer surface proteins. Secondly, the theoretical concerns about autoimmune reactions that could be induced by the vaccine have limited the acceptance of the vaccine by both healthcare providers and people at risk of the disease. A substantial amount of publicity in the lay press has been generated by public concern about the safety of the vaccine despite the reassuring scientific evidence. Thirdly, the currently recommended 0, 1 and 12 month dosing schedule and apparent need for continued boosters, limit the usefulness and effectiveness of the vaccine. Over 100 distinct B. burgdorferi s.l. proteins have been identified and, beyond OspA, a handful of these, including OspC, OspF, decorinbinding protein (Dbp) A, BBK32 and BBK50, have been studied as potential vaccine candidates.

Vaccination Against Lyme Borreliosis

293

OspC vaccines OspC is a major outer surface lipoprotein and immunodominant antigen. This protein is upregulated in the feeding tick and expressed early in infection. Vaccination of outbred mice with recombinant OspC provided protection against infection when the mice were challenged with B. burgdorferi s.s. by tick bite (Gilmore et al., 1996). In addition, passive transfer of OspC antibody resulted in a dose-dependent resolution of fully established infection, arthritis and carditis (Zhong et al., 1999). Unlike OspA antibody, the protective action of OspC antibody appears to take place in the immunized host (Gilmore et al., 1996), which should benefit from an anamnestic response during each exposure. However, significant OspC heterogeneity exists among B. burgdorferi spp. and strains at both antigenic and genetic levels. Over 30 distinct OspC types have been described by restriction fragment length polymorphisms and sequencing (Livey et al., 1995). The N-terminal and C-terminal regions of the mature OspC molecule form relatively conserved domains. A highly variable domain is located in the central portion of the molecule. Comparison of the conserved regions of ospC genes among different isolates indicates that this gene and 16S RNA sequences have evolved at similar rates. However, comparison of the variable region of ospC suggests that ospC diversity has evolved at a greater rate than diversity in other B. burgdorferi s.l. genes (Livey et al., 1995). It has been suggested that immune pressure exerted by an active antibody response to early expression of ospC might play a role in its diversity. However, studies in OspC-hyperimmunized mice do not support this hypothesis; mice infected with clonal B. burgdorferi s.s. or B. afzelii and then immunized with homologous recombinant OspC failed to elicit any sequence changes in ospC over a 6 month period of immune pressure (Hodzic et al., 2000). These findings support the alternative hypothesis that lateral transfer of ospC genes between organisms is the major mechanism for generating diversity of OspC. Mixed infections within tick vectors and mammalian hosts are well documented but proof of a lateral transfer hypothesis awaits tracking of multiple defining markers within such infections. In an attempt to account for antigenic variation within OspC, Baxter developed a pentavalent recombinant OspC vaccine (Livey et al., 1999). In one unpublished, non-randomized trial, 40 seronegative human volunteers were given three doses of either 10, 25 or 50 µg of this vaccine. No serious adverse events were reported. Both reactogenicity and antibody response were more prominent at higher vaccine doses (Livey et al., 1999).

Anti-adhesin vaccines Bacterial adhesion to and colonization of host tissues is a critical step in the establishment of infection (Patti et al., 1994). B. burgdorferi s.l. probably possesses several mechanisms of host and tissue specific adherence. The fibronectin-binding protein, BBK32, and the decorin-binding proteins A and B (DbpA and DbpB)

294

E.B. Hayes and M.E. Schriefer

are several proteins that mediate spirochaete attachment to the extracellular matrix (Guo et al., 1995; Probert and Johnson, 1998) and have been investigated as vaccine candidates and therapeutic agents (Hanson et al., 1998; Fikrig et al., 2000). Expression of bbk32 in the tick is detectable during engorgement and it appears to be upregulated after the spirochaete enters the mammalian host at the cutaneous site of inoculation, and in the heart and spleen. BBK32 immunization of mice, or passive transfer of antiserum, provides partial protection from tick-vectored infection but does not clear established infections. Interestingly, acquisition of infection by ticks and transstadial transmission of spirochaetes in ticks are both reduced in the presence of antibody to BBK32 (Fikrig et al., 2000). Together, these findings suggest that fibronectin-binding protein is important for bacterial adhesion in both the arthropod and mammalian host. Genes for the Borrelia decorin-binding proteins, DbpA and DbpB, have been partially characterized. Immunization of mice with DbpA, as well as passive transfer of serum antibody to DbpA, protects mice from heterologous B. burgdorferi s.s. challenge via needle inoculation. Immunization with DbpB is less protective than immunization with DbpA. Passive transfer of DpbA antibody is therapeutic in mice up to 4 days after infection. Combined immunization with OspA and DbpA has produced protection against 100-fold higher challenges than injection with either protein alone (Hanson et al., 2000). Recent studies have also shown that decorin-deficient mice are more resistant to disease and suggest that this adhesin molecule is important in the spirochaetal colonization of joint tissue (Brown et al., 2001).

Other protein vaccines OspE and OspF often elicit an antibody response during B. burgdorferi s.l. infection of mammals. However, immunization with OspF provides only partial protection against tick-borne or low-dose needle challenge, and immunization with OspE is ineffective (Nguyen et al., 1994). Other antigens that have been tested as vaccine candidates but fail to elicit a protective immune response include the 41 kDa subunit of flagellin, P30, P39, P55 and P83 (Fikrig et al., 1992; Das et al., 1996; Feng et al., 1996; Gilmore et al., 1996).

DNA vaccines Several studies in the early 1990s demonstrated that injection of DNA from pathogenic organisms into experimental animals resulted in humoral and cellular immune responses to antigens encoded by the DNA. DNA vaccines have been shown to be protective against specific viral, bacterial and parasitic infections in experimental animals (Wolff et al., 1990, 1991; Vogel and Sarver, 1995). When mice were immunized with plasmid DNA encoding OspA, they produced

Vaccination Against Lyme Borreliosis

295

OspA-specific antibodies, which inhibited growth of homologous B. burgdorferi s.s. strains in vitro and protected the mice from needle challenge with two heterologous strains (Luke et al., 1997). In contrast to conventional protein antigen immunization, DNA vaccines may be engineered to direct antigen processing and presentation, depending on whether or not specific signal sequences are incorporated into the vaccine (Weiss et al., 1999). The addition of a human tissue–plasminogen–activator signal sequence to DNA encoding OspC enhanced both the humoral and cellular immune responses in vaccinated mice (Weiss et al., 1999). These results suggest that a signal that directs antigen presentation through the endoplasmic reticulum may optimize both the antibody and cellular immune responses to OspCbased plasmid vaccines (Weiss et al., 1999). Other DNA vaccines have been engineered to produce antibody against multiple spirochaetal antigens. In one study a DNA vaccine expressing a fusion product of OspA and OspC was shown to elicit antibody responses to both antigens. However, only OspA antibodies were protective against needle challenge infection and neither antibody response was capable of clearing an established infection (Wallich et al., 2001). No DNA vaccine has yet been shown to protect against tick-transmitted infection of B. burgdorferi s.l.

Anti-tick vaccines Ixodes spp. complex ticks are common vectors of LB, tick-borne encephalitis, ehrlichiosis and babesiosis, and mixed infections can occur within a single tick. Vaccination against any one disease may theoretically lead to a false perception among vaccinees of protection against a greater range of tick-borne diseases and thus minimize the use of other risk-reducing practices or result in delayed diagnosis and treatment. Reduction in vector-tick populations has been shown to have significant impact on LB incidence. However, current control efforts are limited largely to acaricide use and habitat modification. Vaccination against ticks is a relatively new alternative approach. Two targets of this approach are salivary gland antigens that are secreted into the host by the feeding tick (exposed antigens) and others that remain within the tick (concealed antigens) (Willadsen and Jongejan, 1999). The latter do not naturally prime the host’s immune system but are potential targets of antibody ingested by the tick while feeding. Antibody elicited against a recombinant midgut antigen of Boophilus microplus, a tick vector of babesiosis in cattle, has been shown to be effective in controlling infestation by this and other Boophilus spp. as well as providing partial protection against other tick genera (Willadsen and Jongejan, 1999). Other concealed antigen targets might include tick hormones that regulate moulting or reproduction. Repeated feeding by ticks induces a number of immune responses in the host that may lead to inflammation at the attachment site or cause the tick to detach. Identification of exposed antigens from the tick mouth-parts or salivary gland

296

E.B. Hayes and M.E. Schriefer

secretions that elicit a protective rejection of the tick may result in a vaccine that would be boosted by natural exposure.

Summary The development of an efficacious vaccine against LB followed shortly after the identification and characterization of the Borrelia spp. that cause the disease. Immunization with OspA was demonstrated to be protective against infection through the unusual mechanism of eliciting antibodies that presumably destroy or inactivate the spirochaete inside the attached vector tick. In clinical trials a recombinant OspA vaccine with adjuvant has been shown to be nearly 80% efficacious and has an acceptable safety profile. Long-term concerns about the vaccine’s safety centre around evidence that OspA may be a contributing component to possible autoimmune reactions after naturally acquired infection, but there is no evidence that the vaccine directly causes arthritis or other serious adverse events in humans. Rare hypersensitivity reactions to the vaccine have been reported. Despite its limitations, the OspA vaccine is a rare example of a successful transmission-blocking vaccine. A better OspA vaccine might be produced by incorporating antibody epitopes of B. garinii and B. afzelii and removing T cell epitopes that are possibly involved in autoimmune responses. Alternative vaccines could incorporate a combination of different protein antigens, pursue a DNA-directed immune response, or target the vector tick itself in order to provide safe and more complete protection against the variety of Borrelia genospecies that cause LB in North America, Europe and Asia.

References Beran, J., de Clercq, N., Dieussaert, I. and van Hoecke, C. (2000) Reactogenicity and immunogenicity of a Lyme disease vaccine in children 2–5 years old. Clinical Infectious Diseases 31, 1504–1507. Berglund, J., Eitrem, R., Ornstein, K., Lindberg, A., Ringer, A., Elmrud, H., Carlsson, M., Runehagen, A., Svanborg, C. and Norrby, R. (1995) An epidemiologic study of Lyme disease in southern Sweden. New England Journal of Medicine 333, 1319–1327. Brown, E.L., Wooten, R.M., Johnson, B.J., Iozzo, R.V., Smith, A., Dolan, M.C., Guo, B.P., Weis, J.J. and Hook, M. (2001) Resistance to Lyme disease in

decorin-deficient mice. Journal of Clinical Investigation 107, 845–852. Centers for Disease Control and Prevention (1999) Recommendations for the use of Lyme disease vaccine: recommendations of the Advisory Committee on Immunization Practices (ACIP). Morbidity and Mortality Weekly Report 48, 1–24. Das, S., Shraga, D., Gannon, C., Lam, T.T., Feng, S., Brunet, L.R., Telford, S.R., Barthold, S.W., Flavell, R.A. and Fikrig, E. (1996) Characterization of a 30-kDa Borrelia burgdorferi substratebinding protein homologue. Research in Microbiology 147, 739–751. Ding, W., Huang, X., Yang, X., Dunn, J.J.,

Vaccination Against Lyme Borreliosis

Luft, B.J., Koide, S. and Lawson, C.L. (2000) Structural identification of a key protective B-cell epitope in Lyme disease antigen OspA. Journal of Molecular Biology 302, 1153–1164. Edelman, R., Palmer, K., Russ, K.G., Secrest, H.P., Becker, J.A., Bodison, S.A., Perry, J.G., Sills, A.R., Barbour, A.G., Luke, C.J., Hanson, M.S., Stover, C.K., Burlein, J.E., Bansal, G.P., Connor, E.M. and Koenig, S. (1999) Safety and immunogenicity of recombinant Bacille Calmette–Guérin (rBCG) expressing Borrelia burgdorferi outer surface protein A (OspA) lipoprotein in adult volunteers: a candidate Lyme disease vaccine. Vaccine 17, 904–914. Erdile, L.F., Brandt, M.A., Warakomski, D.J., Westrack, G.J., Sadziene, A., Barbour, A.G. and Mays, J.P. (1993) Role of attached lipid in immunogenicity of Borrelia burgdorferi OspA. Infection and Immunity 61, 81–90. Feder, H.M. Jr, Beran, J., van Hoecke, C., Abraham, B., de Clercq, N., Buscarino, C. and Parenti, D.L. (1999) Immunogenicity of a recombinant Borrelia burgdorferi outer surface protein A vaccine against Lyme disease in children. Journal of Pediatrics 135, 575–579. Felz, M.W., Chandler, F.W. Jr, Oliver, J.H. Jr, Rahn, D.W. and Schriefer, M.E. (1999) Solitary erythema migrans in Georgia and South Carolina. Archives of Dermatology 135, 1317–1326. Feng, S., Barthold, S.W., Telford, S.R. III and Fikrig, E. (1996) P55, an immunogenic but nonprotective 55-kilodalton Borrelia burgdorferi protein in murine Lyme disease. Infection and Immunity 64, 363–365. Fikrig, E., Barthold, S.W., Kantor, F.S. and Flavell, R.A. (1990) Protection of mice against the Lyme disease agent by immunizing with recombinant OspA. Science 250, 553–556. Fikrig, E., Barthold, S.W., Marcantonio, N., Deponte, K., Kantor, F.S. and Flavell, R.A. (1992) Roles of OspA,

297

OspB, and flagellin in protective immunity to Lyme borreliosis in laboratory mice. Infection and Immunity 60, 657–661. Fikrig, E., Feng, W., Barthold, S.W., Telford, S.R. III and Flavell, R.A. (2000) Arthropod- and host-specific Borrelia burgdorferi bbk32 expression and the inhibition of spirochete transmission. Journal of Immunology 164, 5344–5351. Gilmore, R.D. Jr, Kappel, K.J., Dolan, M.C., Burkot, T.R. and Johnson, BJ. (1996) Outer surface protein C (OspC), but not P39, is a protective immunogen against a tick-transmitted Borrelia burgdorferi challenge: evidence for a conformational protective epitope in OspC. Infection and Immunity 64, 2234–2239. Gross, D.M., Forsthuber, T., TaryLehmann, M., Etling, C., Ito, K., Nagy, Z.A., Field, J.A., Steere, A.C. and Huber, B.T. (1998) Identification of LFA-1 as a candidate autoantigen in treatment-resistant Lyme arthritis. Science 281, 703–706. Guo, B.P., Norris, S.J., Rosenberg, L.C. and Hook, M. (1995) Adherence of Borrelia burgdorferi to the proteoglycan decorin. Infection and Immunity 63, 3467–3472. Hanson, M.S., Cassatt, D.R., Guo, B.P., Patel, N.K., McCarthy, M.P., Dorward D.W. and Hook, M. (1998) Active and passive immunity against Borrelia burgdorferi decorin binding protein A (DbpA) protects against infection. Infection and Immunity 66, 2143–2153. Hanson, M.S., Patel, N.K., Cassatt, D.R. and Ulbrandt, N.D. (2000) Evidence for vaccine synergy between Borrelia burgdorferi decorin binding protein A and outer surface protein A in the mouse model of Lyme borreliosis. Infection and Immunity 68, 6457–6460. Hayes, E.B., Maupin, G.O., Mount, G.A. and Piesman, J. (1999) Assessing the prevention effectiveness of local Lyme disease control. Journal of Public Health Management and Practice 5, 84–92.

298

E.B. Hayes and M.E. Schriefer

Hodzic, E., Feng, S. and Barthold, S.W. (2000) Stability of Borrelia burgdorferi outer surface protein C under immune selection pressure. Journal of Infectious Diseases 181, 750–753. van Hoecke, C., Comberbach, M., de Grave, D., Desmons, P., Fu, D., Hauser, P., Lebacq, E., Lobet, Y. and Voet, P. (1996) Evaluation of the safety, reactogenicity and immunogenicity of three recombinant outer surface protein (OspA) Lyme vaccines in healthy adults. Vaccine 14, 1620–1626. van Hoecke, C., Lebacq, E., Beran, J. and Parenti, D. (1999) Alternative vaccination schedules (0, 1, and 6 months versus 0, 1, and 12 months) for a recombinant OspA Lyme disease vaccine. Clinical Infectious Diseases 28, 1260–1264. James, A.M., Liveris, D., Wormser, G.P., Schwartz, I., Montecalvo, M.A. and Johnson, B.J. (2001) Borrelia lonestari infection after a bite by an Amblyomma americanum tick. Journal of Infectious Diseases 183, 1810–1814. Kalish, R.A., Leong, J.M. and Steere, A.C. (1993) Association of treatment-resistant chronic Lyme arthritis with HLA-DR4 and antibody reactivity to OspA and OspB of Borrelia burgdorferi. Infection and Immunity 61, 2774–2779. Kalish, R.A., Leong, J.M. and Steere, A.C. (1995) Early and late antibody responses to full-length and truncated constructs of outer surface protein A of Borrelia burgdorferi in Lyme disease. Infection and Immunity 63, 2228–2235. Keller, D., Koster, F.T., Marks, D.H., Hosbach, P., Erdile, L.F. and Mays, J.P. (1994) Safety and immunogenicity of a recombinant outer surface protein A Lyme vaccine. Journal of the American Medical Association 271, 1764–1768. Lathrop, S.L., Ball, R., Haber, P., Mootrey, G.T., Braun, M.M., Shadomy, S.V., Chen, R.T. and Hayes, E.B. (2002) Adverse event reports following vaccination for Lyme disease:

December 1998–July 2000. Vaccine 20, 1603–1608. Livey, I., Gibbs, C.P., Schuster, R. and Dorner, F. (1995) Evidence for lateral transfer and recombination in OspC variation in Lyme disease Borrelia. Molecular Microbiology 18, 257–269. Livey, I., Crowe, B.A., Eder, G., Barrett, N., Gerencer, M., Low-Baselli, A. and Dorner, F. (1999) Clinical studies of a multivalent OspC Lyme disease vaccine. In: VII International Conference on Lyme Borreliosis and other Emerging Tickborne Diseases. Munich, Germany, Abstract S02. Luke, C.J., Carner, K., Liang, X. and Barbour, A.G. (1997) An OspA-based DNA vaccine protects mice against infection with Borrelia burgdorferi. Journal of Infectious Diseases 175, 91–97. Meltzer, M.I., Dennis, D.T. and Orloski, K.A. (1999) The cost effectiveness of vaccinating against Lyme disease. Emerging Infectious Diseases 5, 321–328. Montgomery, R.R., Nathanson, M.H. and Malawista, S.E. (1994) Fc- and non-Fc-mediated phagocytosis of Borrelia burgdorferi by macrophages. Journal of Infectious Diseases 170, 890–893. Nguyen, T.P., Lam, T.T., Barthold, S.W., Telford, S.R. III, Flavell, R.A. and Fikrig, E. (1994) Partial destruction of Borrelia burgdorferi within ticks that engorged on OspE- or OspF-immunized mice. Infection and Immunity 62, 2079–2084. O’Connell, S., Granstrom, M., Gray, J.S. and Stanek, G. (1998) Epidemiology of European Lyme borreliosis. Zentralblatt für Bacteriologie 287, 229–240. Orloski, K.A., Hayes, E.B., Campbell, G.L. and Dennis, D.T. (2000) Surveillance for Lyme disease – United States, 1992–1998. Morbidity and Mortality Weekly Report 49, 1–12. Patti, J.M., Allen, B.L., McGavin, M.J. and Hook, M. (1994) MSCRAMMmediated adherence of microorganisms

Vaccination Against Lyme Borreliosis

to host tissues. Annual Review of Microbiology 48, 585–617. Pavia, C.J., Kissel, V., Bittker, S., Cabello, F. and Levine, S. (1991) Antiborrelial activity of serum from rats injected with the Lyme disease spirochete. Journal of Infectious Diseases 163, 656–659. Probert, W.S. and Johnson, B.J. (1998) Identification of a 47 kDa fibronectinbinding protein expressed by Borrelia burgdorferi isolate B31. Molecular Microbiology 30, 1003–1015. Probert, W.S. and LeFebvre, R.B. (1994) Protection of C3H/HeN mice from challenge with Borrelia burgdorferi through active immunization with OspA, OspB, or OspC, but not with OspD or the 83kilodalton antigen. Infection and Immunity 62, 1920–1926. Schoen, R.T., Meurice, F., Brunet, C.M., Cretella, S., Krause, D.S., Craft, J.E. and Fikrig, E. (1995) Safety and immunogenicity of an outer surface protein A vaccine in participants with previous Lyme disease. Journal of Infectious Diseases 172, 1324–1329. Schoen, R.T., Sikand, V.K., Caldwell, M.C., van Hoecke, C., Gillet, M., Buscarino, C. and Parenti, D.L. (2000) Safety and immunogenicity profile of a recombinant outer-surface protein A Lyme disease vaccine: clinical trial of a 3-dose schedule at 0, 1, and 2 months. Clinical Therapeutics 22, 315–325. Shadick, M.A., Liang, M.H., Phillips, C.B., Fossel, K. and Kuntz, K.M. (2001) The cost-effectiveness of vaccination against Lyme disease. Archives of International Medicine 161, 554–561. Sigal, L.H., Zahradnik, J.M., Lavin, P., Patella, S.J., Bryant, G., Haselby, R., Hilton, E., Kunkel, M., Adler-Klein, D., Doherty, T., Evans, J., Molloy, P.J., Seidner, A.L., Sabetta, J.R., Simon, H.J., Klempner, M.S., Mays, J., Marks, D. and Malawista, S.E. (1998) A vaccine consisting of recombinant Borrelia burgdorferi outer-surface protein A to pre-

299

vent Lyme disease. New England Journal of Medicine 339, 216–222. Sikand, V.K., Halsey, N., Krause, P.J., Sood, S.K., Geller, R., van Hoecke, C., Buscarino, C. and Parenti, D. (2001) Safety and immunogenicity of a recombinant Borrelia burgdorferi outer surface protein A vaccine against Lyme disease in healthy children and adolescents: a randomized controlled trial. Pediatrics 108, 123–128. de Silva, A.M., Telford, S.R. III, Brunet, L.R., Barthold, S.W. and Fikrig, E. (1996) Borrelia burgdorferi OspA is an arthropod-specific transmission-blocking Lyme disease vaccine. Journal of Experimental Medicine 183, 271–275. Smith, P.F., Benach, J.L., White, D.J., Stroup, D.F. and Morse, D.L. (1988) Occupational risk of Lyme disease in endemic areas of New York state. Annals of the New York Academy of Sciences 539, 289–301. Steere, A.C., Sikand, V.K., Meurice, F., Parenti, D.L., Fikrig, E., Schoen, R.T., Nowakowski, J., Schmid, C.H., Laukamp, S., Buscarino, C., Krause, D.S. and the Lyme Disease Vaccine Study Group (1998) Vaccination against Lyme disease with recombinant Borrelia burgdorferi outer surface lipoprotein A with adjuvant. New England Journal of Medicine 339, 209–215. Telford, S.R. and Fikrig, E. (1995) Progress towards a vaccine for Lyme disease. Clinical Immunotherapy 4, 49–60. Vogel, F.R. and Sarver, N. (1995) Nucleic acid vaccines. Clinical Microbiology Reviews 8, 406–410. Wallich, R., Siebers, A., Jahraus, O., Brenner, C., Stehle, T. and Simon, M.M. (2001) DNA vaccines expressing a fusion product of outer surface proteins A and C from Borrelia burgdorferi induce protective antibodies suitable for prophylaxis but not for resolution of Lyme disease. Infection and Immunity 69, 2130–2136. Weiss, R., Durnberger, J., Mostbock, S.,

300

E.B. Hayes and M.E. Schriefer

Scheiblhofer, S., Hartl, A., Breitenbach, M., Strasser, P., Dorner, F., Livey, I., Crowe, B. and Thalhamer, J. (1999) Improvement of the immune response against plasmid DNA encoding OspC of Borrelia by an ER-targeting leader sequence. Vaccine 18, 815–824. Willadsen, P. and Jongejan, F. (1999) Immunology of the tick–host interaction and the control of ticks and tick-borne diseases. Parasitology Today 15, 258–262. Wolff, J.A., Malone, R.W., Williams, P., Chong, W., Acsadi, G., Jani, A. and Felgner, P.L. (1990) Direct gene transfer into mouse muscle in vivo. Science 247, 1465–1468. Wolff, J.A., Williams, P., Acsadi, G., Jiao,

S., Jani, A. and Chong, W. (1991) Conditions affecting direct gene transfer into rodent muscle in vivo. Biotechniques 11, 474–485. Wormser, G.P. (1995) Prospects for a vaccine to prevent Lyme disease in humans. Clinical Infectious Diseases 21, 1267–1274. Zhong, W., Gern, L., Stehle, T., Museteanu, C., Kramer, M., Wallich, R. and Simon, M.M. (1999) Resolution of experimental and tick-borne Borrelia burgdorferi infection in mice by passive, but not active immunization using recombinant OspC. European Journal of Immunology 29, 946–957.

Environmental Management for Lyme Borreliosis Control

12

Kirby C. Stafford1 and Uriel Kitron2 1

Connecticut Agricultural Experiment Station, PO Box 1106, New Haven, CT 06540, USA; 2Department of Veterinary Pathobiology, College of Veterinary Medicine, University of Illinois, 2001 S. Lincoln, Urbana, IL 61802, USA

Introduction Ticks are obligate haematophagous ectoparasites of humans, livestock and wildlife, associated with many zoonotic disease pathogens. Tick management has historically been more of a concern in the veterinary realm, where control options such as rotational grazing, dipping and quarantine are also available. Control of human tick-borne diseases may target the vector or the pathogen (Jaenson et al., 1991; Tatchell, 1992; Wilson and Deblinger, 1993; Schmidtmann, 1994; Sonenshine, 1994) with the purpose of reducing tick–human contact or reducing the prevalence of infection in the tick vector and/or the vertebrate hosts. Strategies against the vector include personal protection from tick bite and approaches that reduce tick introduction into tick-free areas, tick abundance or tick reproduction. The pathogen is targeted through vaccination or, indirectly, through methods to reduce the prevalence of infection in the vector or reservoir host. Methods that have been explored to control ticks include biological control, area chemical applications, acaricidal treatment of small and large mammalian hosts, habitat modification, reduction in host availability and integrated pest management (IPM) strategies. Other approaches, such as an anti-tick vaccine, remain experimental or speculative in terms of human disease. The commercial availability of a recombinant vaccine against Boophilus microplus is an important new step in the control of ticks on livestock (Willadsen et al., 1995). Four tick species of the genus Ixodes that readily feed on humans, Ixodes scapularis, Ixodes pacificus, Ixodes ricinus and Ixodes persulcatus, are competent vectors of the Borrelia spp. that cause Lyme borreliosis (LB). They are distributed over broad areas of North America, Europe and Asia, placing many people at risk of not only LB, but also other tick-borne infections carried by these ticks. Control © 2002 CAB International. Lyme Borreliosis: Biology, Epidemiology and Control (eds J. Gray, O. Kahl, R.S. Lane and G. Stanek)

301

302

K.C. Stafford and U. Kitron

of LB has relied heavily on personal protection behaviours, antibiotic and supportive therapy after infection, and to a lesser extent the application of acaricides to tick habitats. The environmental management of tick populations and reduction in risk of the diseases they transmit requires knowledge of tick ecology and the unique habitats of each tick species, the transmission dynamics of LB (which may vary by Borrelia genospecies) and a proper assessment of risk. The wide geographical distribution, abundance and host range of these ticks, and the diversity of known reservoirs for Borrelia burgdorferi sensu lato (s.l.) and other genospecies complicate the management or control of tick-borne disease. Tick control over large areas was the leading strategy for the prevention of tick-borne encephalitis (TBE) in Russia from about 1950 through to the 1970s (Uspensky, 1996, 1999). The use of DDT in a large-scale campaign successfully suppressed I. persulcatus in high-risk areas of taiga forest resulting in substantial reductions in human contacts with ticks and decreased incidence of TBE. With the gradual reduction and cessation of tick control programmes in Russia, TBE and the more recently recognized LB has increased (Korenberg, 1994). Although acaricides are still an important element of tick management, environmental and toxicological safety concerns have restricted the availability and wide-scale use of these chemicals. Nevertheless, acaricides are still commonly used for shortterm control of I. scapularis around private residences and other localized highrisk areas (Stafford, 1997). While a variety of effective acaricides, mainly pyrethroids, are still available, there is a need for newer, less toxic alternatives and the integration of these materials with other methodologies. Targeting the acaricides to tick hosts is one approach. As the prevalence of LB continues to increase, environmentally acceptable approaches are needed to manage tick populations and reduce the risk of acquiring disease. Human tick-borne illnesses require an ecologically based integrated management approach based on an assessment of risk. The ecology of the various Ixodes spp., which feed on a diverse range of wildlife and are broadly distributed geographically, is complex. No single methodology will address the cycle of tick vector, vertebrate reservoirs, spirochaetes and exposure of the susceptible human population, and focused efforts to reduce the incidence of LB are few. In this chapter, we address the assessment of risk for LB, current strategies in managing tick populations, approaches that are experimental, and future prospects and strategies for tick control and the management of human disease.

Habitat Assessment and Risk While individual ticks operate in the microenvironment and encounter only one host at each stage, risk assessment and control strategies to reduce the probability of tick–human encounters and the transmission of tick-borne diseases need to be considered over a range of spatial scales (Cortinas et al., 2002). Thus, potential individual human hosts apply measures of personal protection and targeted acaricides are directed towards ticks infesting individual reservoir hosts. In

Environmental Management for LB Control

303

contrast, landscape and host management and area application of acaricides operate on a wider geographical range and affect a large number of hosts and vectors. Tick habitat can be viewed on the micro-, meso- and macroscales ranging from the ground level microenvironment to the country- or continent-wide epidemiological picture (Kitron, 1998, 2000; Cortinas et al., 2002). Satellite imagery and geographical information systems (GIS) have been used to map tick and LB risk distribution (Daniel and Kolar, 1990; Nicholson and Mather, 1996; Kitron and Kazmierczak, 1997; Daniel et al., 1998), and can be used as surveillance tools to highlight risk areas and periods and to determine recommendations for application of control measures. Risk maps for transmission of LB (and other diseases transmitted by Ixodes spp. ticks) have been developed primarily based on habitat suitability for tick distribution in the USA (Kitron et al., 1992; Glass et al., 1994, 1995; Kitron and Kazmierczak, 1997), although the zooprophylactic effect of lizards and the effect of deer exclusion have also been considered (Fish and Howard, 1999). In Europe, where the variability in B. burgdorferi s.l. genotypes is well documented, tick distribution has also been associated with environmental features (Daniel and Kolar, 1990; Merler et al., 1996; Gray et al., 1998), but the role of alternative hosts in transmission of the various B. burgdorferi subspecies has also been considered and contrasted with the requirements for transmission of tick-borne encephalitis (Randolph, 2000; Randolph and Rogers, 2000; Randolph et al., 2000). On a very fine scale, Dister et al. (1997) used Landsat Thematic Mapper (TM) remotely sensed data to categorize residential properties in Westchester County, New York, as no, low or high risk based on vegetation structure, moisture and abundance. Thus, residential mapping was used to determine habitat suitability for ticks and risk of transmission of LB within communities. Such an analysis can be used to derive landscape maps that can be used to target intervention techniques (e.g. deer exclusion, area insecticide application to specific properties). TM data were also used by Daniel et al. (1998) to detect micro-foci in Bohemia, Czech Republic. In Maryland and Rhode Island, USA, risks at county and state level were assessed using satellite data and GIS (Glass et al., 1995; Nicholson and Mather, 1996). In Wisconsin, the distribution of human LB cases on the county level was associated with forest cover (Kitron and Kazmierczak, 1997). A risk map for Wisconsin and Illinois was developed based on canine serology and tick distribution associated with a range of environmental features (Cortinas et al., 2002; Guerra et al., 2001a,b) and can be used to target surveillance efforts and control priorities. On a continental scale, Fish and Howard (1999) provided a coarse continental map for LB risk in the USA that can be used to determine whether vaccination should be recommended or even considered. In Europe, a collaborative effort provided descriptions of LB habitats from 16 countries and these were used to associate high-risk situations with heterogeneous deciduous forest, recreational activities and diverse fauna (Gray et al., 1998). In a recent review, Randolph (2000) provided a comprehensive summary and

304

K.C. Stafford and U. Kitron

analysis of different approaches to identify tick habitat, and highlighted the biological basis underlying critical satellite signals. Randolph et al. (2000) used coarse resolution NOAA-AVHRR (National Oceanic and Atmospheric Administration Advanced Very High Resolution Radiometer) satellite imagery and considered both NDVI (Normalized Difference Vegetation Index) and LST (Land Surface Temperature). They processed monthly images using temporal Fourier analysis, and contrasted the risk of LB transmission with the much more restricted geographical risk of TBE transmission. Much of the biological knowledge necessary to successfully apply satellite data to understand and produce risk maps was based on field and laboratory studies that determined the necessary condition for transmission of the various tick-borne pathogens. In contrast, efforts to produce continental risk maps, based on meta-analysis of published data (Estrada-Peña, 1998, 2001) and on cokriging of simple habitat features (e.g. NDVI), can result in maps that are discordant with known distribution of ticks (e.g. habitat suitability versus known distribution in Wisconsin, USA, and in Britain and Ireland). In Europe, in particular, the biological diversity of LB transmission systems (due to genetic diversity of the spirochaete and variations in reservoir competence) renders predictive risk maps less precise. The ability to incorporate large amounts of environmental data (some of them through remote sensing) with biological information allows for effective targeting of control measures on a range of spatial scales. Accurate maps of tick distribution and disease risk are based on good understanding of tick biology and pathogen transmission, as are effective control strategies. In particular, spatial tools for habitat assessment can play an important role in environmental management and integration of control measures where consideration of habitat heterogeneity on multiple spatial scales is necessary for successful implementation.

Measures for Control of Ixodes spp. Ticks and Lyme Borreliosis Control measures can be divided into measures applied on an individual basis (personal protection and vaccination) and measures that target the tick population (on a range of spatial and temporal scales) (Table 12.1). In this chapter, we emphasize the latter measures, which are applied in the environment. After a brief review of measures for prevention of tick bites, treatment and vaccination (including host immunity and anti-tick vaccines), we concentrate on management of landscape and hosts and area application of acaricides. We also discuss the actual and potential use of biological control agents and pheromones, whose application is part of an environmentally based strategy, even though they may target individual ticks. We conclude by discussing the use of various strategies in an IPM programme and the future of tick and LB control. Personal protection and behaviours to prevent or reduce tick bite are the first

Environmental Management for LB Control

305

Table 12.1. Existing and potential management strategies for the control of Lyme borreliosis. These approaches may reduce exposure to the tick vector, target the tick vector or target the pathogen. Personal protection

Immunologically based approaches Landscape management

Management host abundance and availability Area application of acaricides Host-targeted acaricides

Biological control

Pheromones Integrated pest management (IPM)

Tick-bite prevention to minimize exposure: protective clothing, use of repellents. Tick removal and tick-bite prophylaxis: full examination of the body and prompt removal of ticks, prophylactic treatment with antimicrobials. Vaccination against Borrelia burgdorferi. Early diagnosis and treatment of infection. Research into immunological interactions of the tick–host–pathogen interface. Development of anti-tick vaccines. Vegetative modifications (mowing, clearing leaves, xeric substrates) and burning to render the environment less suitable for tick survival and for tick hosts. Use of deer-resistant plants to reduce host visitation. Exclusion of hosts from areas by fencing. Host reduction by management, elimination or alteration of the environment. Treatment of white-tailed deer with an acaricide through passive topical application devices. Treatment of white-footed mice with permethrintreated cotton or fipronil-based bait boxes. Parasitoids: potential of encrytid wasp Ixodiphagus hookeri for tick management. Pathogens: potential of fungal, nematode and other pathogens for tick management. Predators: role of predators in reducing tick abundance. Use of pheromone-based lures and decoys with acaricides for tick control, either on or off host. Concepts of IPM related to ticks and vectorborne disease and application of IPM strategies. Use of computer models to simulate various strategies for the prevention of Lyme borreliosis.

306

K.C. Stafford and U. Kitron

line of defence in the prevention of LB. These include avoidance and reduction of time spent in tick-infested habitats, using protective clothing and tick repellents, checking the entire body for ticks and promptly removing attached ticks before transmission of borrelial spirochaetes can occur. However, the increasing incidence of disease and surveys in the northeastern USA (Cartter et al., 1989; Brown et al., 1992; Shadick et al., 1997; Phillips et al., 2001) suggest that few people practise these measures consistently enough for them to be effective. Despite the efficacy of tick repellents, particularly with N, N-diethyl-3-methylbenzamide (DEET) applied to skin and permethrin applied to clothing, they are grossly underused. Preventive measures are often considered inconvenient; those who believe their chance of acquiring LB is high are more likely to take precautions and frequent detection of ticks appears to be a major incentive for engaging in prevention behaviours. When the tick feeds, spirochaetes residing in the gut of an infected tick multiply; some migrate through the gut into the haemolymph, disseminate to the salivary glands and are introduced into the host during the later stages of feeding (Ribeiro et al., 1987; Piesman, 1995; Gern et al., 1996). The risk of infection is low during the first 24–48 h of attachment by I. scapularis (des Vignes et al., 2001; Ohnishi et al., 2001) and I. ricinus (Kahl et al., 1998) and increases after attachment for 48 h or more. Consequently timely tick removal and tick-bite prevention are effective control measures. However, the prophylactic use of antibiotics following a tick bite has not generally been recommended in the USA (Warshafsky et al., 1996), as the chance of infection from a known tick bite with I. scapularis appears low (< 5%, about equal to the chance of a reaction to the antibiotic) and the majority of patients diagnosed with LB do not detect the tick (Costello et al., 1989; Shapiro et al., 1992). Nevertheless, a large proportion (55%) of patients with tick bite apparently receive antibiotic therapy (Fix et al., 1998). Risk is higher with engorged ticks and, although tick testing is not a reliable predictor of infection, prophylactic treatment may be justified in such cases (Sood et al., 1997). Transmission rates with I. ricinus, and particularly with adult I. persulcatus, may be higher and treatment may reduce the risk of LB (Alekseev et al., 1996; Maiwald et al., 1998). Although not strictly a prevention strategy, early diagnosis and treatment are important in the management of LB. Early treatment has an excellent prognosis and can prevent the development of later stages of disease (Wormser et al., 2000). The use of recombinant outer surface protein A (OspA) vaccines was found to prevent human LB in large-scale safety and efficacy trials (Sigal et al., 1998; Steere et al., 1998), and one vaccine (LYMErixTM, GlaxoSmithKline), consisting of lipidated OspA with an adjuvant, was approved in the USA in December 1998 (Centers for Disease Control and Prevention, 1999a). European isolates of borrelia show great variation in their outer surface proteins and may express OspA poorly or not at all. Vaccines that utilize a complex of other candidate antigens, such as OspC or a cocktail of various OspAs, may be needed for an effective European borreliosis vaccine (Gern et al., 1997). Recommendations for use of the vaccine are based on the risk of exposure to

Environmental Management for LB Control

307

infected ticks (Centers for Disease Control and Prevention, 1999b). The economic benefit of vaccinating against LB is dependent mainly on the probability of contracting the disease, followed by the cost of treating cases and the probability of early detection and treatment of LB (Meltzer et al., 1999). In modelling the effectiveness of various prevention measures, Hayes et al. (1999) found that use of the vaccine was among the more effective options in preventing cases in the simulated scenarios. For example, vaccination would prevent 148 cases of LB over a 4 year period in a hypothetical population of 10,000 individuals, assuming 65% of the population were targeted, 74% of those were actually vaccinated and vaccine efficacy was 60% with two doses and 90% for the first year after three doses. The future role that an LB vaccine could play, as a component of an integrated programme for the prevention of LB, is unclear as the first LB vaccine (LYMErixTM, GlaxoSmithKline) was withdrawn from the market by the manufacturer in February 2002 due to a lack of sales. Future improvements should focus on developing second-generation vaccines, possibly through antigens expressed in the human host. Vaccination is discussed in depth by Hayes and Schriefer in Chapter 11. Another promising approach in the control of ticks is immunologically based, focusing on the development of anti-tick vaccines (Wikel et al., 1996; Willadsen and Jongejan, 1999). Ixodid ticks attach and feed on their hosts for many days, introducing through the saliva a variety of pharmacologically active compounds that aid the feeding process, improve engorgement success, suppress innate and acquired immunity of the host and affect pathogen transmission (reviewed by Bowman et al., 1997). These elements also function as antigens that stimulate the host immune response. There is a balance between immune-mediated resistance by the host and strategies employed by the tick to circumvent these mechanisms. Vaccination with salivary gland-derived or other antigens can induce resistance similar to naturally acquired resistance. An anti-tick feeding vaccine could inhibit tick-feeding success and potentially reduce transmission of multiple pathogens from that vector. Vaccination with a commercial recombinant vaccine against ticks has proved effective as a component in the control of cattle ticks and of bovine babesiosis (Willadsen et al., 1995) and offers potential for reducing transmission of tickborne infections in reservoir hosts or the prevention of disease in humans. Research currently focuses on a better understanding of the biological basis by which ticks modulate host immune responses and finding tick-derived proteins that could be used for an anti-tick vaccine. A similar approach, vaccination or treatment of wild animal reservoirs with antimicrobials, may also have potential to reduce the transmission of B. burgdorferi s.l. and subsequently the risk of human infection. The vaccination of B. burgdorferi-infected Peromyscus leucopus (white-footed mouse) with recombinant OspA reduced transmission of spirochaetes to ticks fed upon vaccinated animals (Taso et al., 2001).

308

K.C. Stafford and U. Kitron

Landscape management Modification of the tick microenvironment to reduce relative humidity and soil moisture has long been considered effective in reducing the abundance of ticks. Free-living ixodid ticks spend the majority of their life off the host in the cover vegetation, leaf litter, duff or other accumulated vegetative debris containing a suitable microclimate for survival and search for hosts. Survival is a function of maintaining a balance between energy use and body water homoeostasis, which is affected by permeability of the cuticle, water mass, sensitivity to desiccation and ability to absorb atmospheric water vapour (Needham and Teel, 1986; Kahl and Knülle, 1988). Therefore, humidity is crucial to the survival of many ticks, including I. ricinus and I. scapularis: with their small water mass and relatively permeable cuticle, they cannot survive as long as some other mesic tick species (Lees, 1946; Yoder and Spielman, 1992; Stafford, 1994). Modifications to the landscape and alterations to the vegetation to produce an unfavourable microclimate for tick oviposition, egg hatch, moulting and host seeking, or for vertebrate hosts may be particularly effective against the hydrophilic subadult stages of Ixodes spp. The influence of habitat structure on populations of ticks and tick hosts determines in turn the risk of human exposure to LB. Adler et al. (1992), for example, found that larval burdens on whitefooted mice were positively associated with woody vegetation density and negatively with grassy and herbaceous vegetation. Nymphal burdens were also negatively correlated with herbaceous vegetation, but less so than the larvae as nymphs are transported and distributed (as engorged larvae) by birds and mammals. Mowed lawns with relatively little shade have fewer blacklegged tick nymphs despite high tick densities in adjacent ornamental ground cover and wooded areas, and a barrier of xeric landscaping materials at the lawn–ecotone interface could reduce residential exposure (Maupin et al., 1991; Stafford and Magnarelli, 1993). A number of studies have examined the impact of vegetation management on ticks. Mechanical clearing of the vegetation, clearing with herbicides, removal of leaf litter and controlled burns have been used to modify the vegetative environment. The mechanical or chemical clearing of vegetation has been shown in numerous studies to suppress populations of the lone star tick, Amblyomma americanum (e.g. Mount, 1981; Barnard, 1986). Vegetation management has also been shown to affect I. scapularis. Mowing and burning produced a dramatic, but temporary decline in adult tick abundance (Wilson, 1986), while removal of leaf litter in wooded residential plots or at lawn perimeters reduced the abundance of I. scapularis nymphs significantly (Schulze et al., 1995; K. Stafford, unpublished data). Since untreated landscape stones and pinebark woodchips were found to deter the movement of I. scapularis nymphs in the laboratory (Patrican and Allan, 1995b), a field study on residential lawns found that the use of a woodchip border (Fig. 12.1) reduced the number of I. scapularis nymphs on the lawn perimeter by approximately 50% on average (K. Stafford, unpublished data). Fire has been used extensively as a tool for forest and range management. It

Environmental Management for LB Control

309

Fig. 12.1. Landscape management using a woodchip barrier at the woodland–residential lawn interface to reduce movement of ticks into the lawn; a perimeter application of an acaricide can increase the efficacy of the barrier treatment. (Photograph by K. Stafford.)

was also one of the earliest tools advocated to control ticks, and grass burns were used in the attempt to control the Rocky Mountain wood tick and spotted fever (Fricks, 1915). Controlled burns were reported to reduce the abundance of lone star ticks (Hoch et al., 1972). Significantly lower numbers of I. scapularis were found in pine palmetto flatwoods that had been burned 1–2 years prior to sampling than an area last burned 14 years previously (Rogers, 1955). Stafford (1998) found that a single, light to moderate controlled spring burn in a northeastern deciduous forest reduced nymphal ticks by 74%. A moderate to severe burn resulted in a reduction of nymphs by 97% in comparison with an adjacent unburned woodland. However, the effect appeared temporary judging by the comparable number of adult I. scapularis recovered in the burned and unburned tracts the following autumn. Similar results were obtained in a state park in northern Illinois (Fig. 12.2; C. Jones and U. Kitron, unpublished data). Vegetation burns on Great Island, Massachusetts, reduced adult I. scapularis by 70 and 80% following the burns (Wilson, 1986). Again, the effect was temporary and no effect on adult tick abundance could be detected 1.5 years later. In a single burn on Shelter Island, New York, 49% fewer nymphs were observed compared with an unburned tract, but the fire did not penetrate the humus layer (Mather et al., 1993). The impact on larval and adult ticks later in the season was not reported. The potential use of the destruction of vegetation by controlled burns is limited as the main areas for risk of LB in the USA are in relatively densely populated residential areas and burning woodlands can improve deer browsing and the density of Peromyscus spp., which would increase host densities.

310

K.C. Stafford and U. Kitron

Fig. 12.2. Controlled burning in Castle Rock State Park in northern Illinois, USA. Although intended for vegetation control and landscape management, the effects of the burn on tick and small mammal densities were monitored as part of a CDC-sponsored tick control study. (Photograph by C. Jones and U. Kitron.)

Host management, reduction and exclusion Members of the I. ricinus species complex associated with LB are three-host ticks with a broad host range, attacking many avian and mammalian hosts (see Chapters 6–9 on the ecology of B. burgdorferi s.l.). Many of these animals are reservoirs for B. burgdorferi, and white-footed mice, P. leucopus, in particular, are important hosts for the spirochaete where I. scapularis is found. White-tailed deer (Odocoileus virginianus) are the preferred host for the adult stage of I. scapularis and are crucial for the reproduction of this tick, but not for the transmission of B. burgdorferi. These vertebrate hosts have been closely linked to the abundance and distribution of I. scapularis and indirectly to the rising incidence of LB. Management of tick hosts has been an important part of controlling cattle ticks (Boophilus spp.) by denying a primary host to free-living larvae of these onehost ticks through pasture spelling, sometimes conducted with rotational grazing (Wilkinson, 1957). Pasture spelling has also been effective against I. ricinus by denying this tick access to sheep (Randolph and Steele, 1985). The presence of B. burgdorferi and Babesia microti, an agent of human babesiosis, and their vectorticks was examined on two islands inhabited by deer and mice and four islands without deer in Rhode Island (Anderson et al., 1987). Both pathogens were isolated from tick-infested white-footed mice and meadow voles (Microtus pennsylvanicus) on the two deer-inhabited islands. Neither of the pathogens nor

Environmental Management for LB Control

311

the blacklegged tick was detected in mice on the four islands that had no deer. This suggested that, in the absence of deer, other mammalian hosts could not sustain the tick in sufficient numbers to transmit the LB or babesiosis agents. However, a population of the bird-feeding Ixodes uriae and Borrelia garinii were maintained on a Baltic island and on the Faeroe Islands where marine birds were the only vertebrate host, demonstrating that deer and/or other mammals are not a prerequisite for completion of the life cycle and transmission of B. burgdorferi s.l. (Olsen et al., 1993; Gylfe et al., 1999). The exclusion or reduction of white-tailed deer could potentially reduce vector abundance and affect the dynamics of disease transmission. When engorged female ticks drop off deer, they seek protected microclimates under the leaf litter or grass to lay their eggs. Hatching larvae actively disperse only a few metres from the egg mass (Lees and Milne, 1951; Stafford, 1992a), relying upon a passing host for survival and dispersion. Therefore, the exclusion of deer, with the corresponding reduction in egg-laying female ticks, would be expected to have the greatest effect on the number of larvae. Indeed, the exclusion of deer by electric fencing from two relatively small areas (≈ 3.5 and 7.4 ha) in Connecticut was shown to reduce the abundance of I. scapularis, particularly larvae, within the fenced areas (Stafford, 1993). A seven-wire, slanted, high-tensile electric deer fence resulted in reductions in larvae of 88.5–97.8% and nymphs of 47.4–55.8% within the fence compared with outside the deer exclosures. Daniels et al. (1993) reported 83% fewer nymphs and 90% fewer larvae inside five exclosures with traditional fencing. These reductions were similar to those obtained for A. americanum larvae (98%) and nymphs (53%) in a 2.4 ha deer exclosure at the Land Between the Lakes recreational area in Tennessee (Bloemer et al., 1986, 1990). Similarly, the exclusion of fallow deer, Dama dama, by a permanent fence resulted in significantly lower densities of I. ricinus nymphs compared with an adjacent forested area inhabited by deer in Ireland (Gray et al., 1992). The size of the exclusion area is a primary determinant of the impact of exclusion on the tick population. Small and medium-sized mammals and birds are important to the passive dispersal of immature I. scapularis and these animals may reintroduce ticks within a deer exclosure, particularly near the edges (Daniels and Fish, 1995). Therefore, deer fencing alone can significantly reduce, but not eliminate, nymphal and adult tick populations and the risk of LB. There may also be local restrictions on the use, size and type of fencing permitted. Electric fences require regular maintenance, suitable terrain and proper grounding, and may be inappropriate in areas with a potential for frequent human contact (Ellingwood and Caturano, 1988). Another strategy for individual residences is the use of repellents and browse-resistant plantings to discourage deer. However, deer repellents may be ignored and less palatable vegetation browsed when alternative food is scarce or deer densities are high. The incremental removal of deer has been shown to reduce tick abundance (Deblinger et al., 1993; K. Stafford, unpublished data). A gradual reduction in deer densities over a 7 year period at a 567 ha site in Ipswich, Massachusetts, reduced the abundance of I. scapularis larvae and nymphs on white-footed mice

312

K.C. Stafford and U. Kitron

by roughly one-half, although the density of female ticks on deer increased fourto sixfold during the 7 year reduction of deer at Crane’s Beach (Deblinger et al., 1993). Similarly, the reduction of deer from 97.3 km–2 to 13.1 km–2 over a 7 year period on a 176 ha forested, fenced property in Connecticut, resulted in an almost fivefold reduction in nymphal tick abundance (K. Stafford, unpublished data). The virtual elimination of deer from Great Island, a 240 ha coastal peninsula near Cape Cod, Massachusetts, produced a dramatic reduction in the subadult population of this tick (Wilson et al., 1988a,b), although adult tick abundance initially seemed to increase after deer became scarce. This is due to the 2 year life cycle of the tick and the fact that a reduction in deer density (or exclusion), in the short term, may increase apparent host-seeking activity of adult ticks, as fewer deer will remove fewer adult ticks from the environment (Ginsberg and Zhioua, 1999). Adult tick populations eventually decline with the removal of deer. While deer removal has not eliminated the tick or interrupted the enzootic cycle on Great Island (adult ticks still occur on dogs, raccoons and foxes), the elimination of deer has apparently reduced or prevented human cases of LB on Great Island and observations suggest a deer density of 3 km–2 will reduce tick densities sufficiently to interrupt tick–human transmission of LB (Telford, 1993). The computer simulation model LYMESIM was developed to assess the population dynamics of I. scapularis (Mount et al., 1997a). The model was used to examine the relationship between the density of white-footed mice and whitetailed deer in coastal Connecticut and the density and B. burgdorferi infection rates in the tick. Tick density thresholds for the maintenance of B. burgdorferi were also examined. There was a linear relationship between the density of free-living nymphal I. scapularis and mice, with tick density and infection rate influenced by the density of alternative hosts. In contrast, there was a curvilinear relationship between deer and tick density, with ticks disappearing when deer density was reduced to zero. The infection rate in nymphs increased with moderate decreases in deer density. In the short term, the reduction in numbers of a major host species may result in increased feeding on alternative hosts, minimizing the impact on the tick population and the prevalence of infection in the tick population. Actual reduction (or exclusion) of deer has resulted in a notable increase in the prevalence of B. burgdorferi in host-seeking nymphs (Gray et al., 1992; K. Stafford, unpublished data). In the simulation, however, infection rates dropped with less than 0.75 deer km–2 and went to zero at 0.3 deer km–2. There was a reduced density of ticks with each decrease in deer density. In the model, a density of I. scapularis below a threshold of 87 nymphs ha–1 does not maintain B. burgdorferi in the ecosystem. Mount et al. (1997b) extended the model to simulations of several management strategies on tick density, including the impact of reduced deer density on the abundance of I. scapularis infected with B. burgdorferi. In a 10 year simulation, an initial deer density of 25 animals km–2 was reduced to 7.5, 2.5 and 0.25 deer km–2 (70, 90 and 99%, respectively). With only 0.25 deer km–2, infected nymphs declined by 74% by year 3 and 98% by year 10.

Environmental Management for LB Control

313

The removal of most of the deer or complete elimination of the animal presents some severe logistical and political obstacles. Deer management has traditionally relied on regulated seasonal or controlled hunting and implementation can be controversial, particularly on public lands (Kilpatrick et al., 1997). With the exception of some islands, it would be difficult to reduce deer sufficiently in most environments to affect tick populations or the dynamics of disease transmission. Conversely, unmanaged deer populations can result in superabundant tick populations as evident on several islands and other isolated tracts along the New England coast. Within a community, the reduction or elimination of deer requires the approval of the majority of community residents. The successful implementation of a deer reduction programme will require clear communication to the public and press of the biological basis and objectives for any wildlife reduction (i.e. protection of biodiversity, reduction in vehicle collisions or tickassociated disease), the support of stakeholders (local residents, politicians and the public) and good estimates of the size of the deer population (DeNicola et al., 2000). However, the need for such a drastic reduction in deer population to achieve satisfactory control levels is likely to render this strategy unrealistic or unacceptable in many, if not most, areas.

Area application of acaricides Although the large-scale application of acaricides against ticks is precluded by current environmental and safety concerns, these chemicals have been shown to provide effective short-term suppression of I. scapularis in applications targeting primarily the nymphal stage associated with most cases of LB. In early trials against I. scapularis, autumn application of carbaryl or diazinon was shown to reduce adult I. scapularis populations in an oak-dominated forest for 1 year and a single application of carbaryl to lawns and adjacent grassy and woodland areas of residential properties provided 90.8% control of I. scapularis nymphs for almost 2 months (Schulze et al., 1987; Stafford, 1991). Curran et al. (1993) reported that various concentrations of liquid and granular formulations of carbaryl and chlorpyrifos, and a single spray application of cyfluthrin provided 63.4–85.7%, 84.9–95.6% and 93.1% reductions, respectively, in the density of nymphal I. scapularis for as long as 41 days after application. Granular formulations of diazinon, chlorpyrifos and carbaryl also provide effective control of I. scapularis (Schulze et al., 1991, 1994, 2000). Similar results with chlorpyrifos and carbaryl were reported for adult I. pacificus in northern California foothills (Monsen et al., 1999). The effectiveness of materials is dependent on contact with the litter layer, depth of the leaf litter and sufficient pressure to penetrate vegetation and the leaf litter (Schulze and Harrison, 1995). Chemical control as a strategy to reduce tick abundance has principally been applied in residential settings in the USA. In a survey of licensed commercial pesticide applicators in Connecticut, treatment of residential properties comprised a larger proportion (90.1%) of tick control services than commercial

314

K.C. Stafford and U. Kitron

properties (9.9%) (Stafford, 1997). The principal acaricides used for the control of I. scapularis were cyfluthrin, chlorpyrifos and carbaryl. A more recent survey in the heavily populated southeastern portion of the state found that the pyrethroid cyfluthrin was used by over half of the commercial applicators (K. Stafford, unpublished data). It is unknown how many homeowners apply acaricides to control ticks, but only 22.5% of 402 respondents in an LB survey of two Connecticut towns indicated that they had used acaricides and of those 16.9% indicated someone in their household had applied the chemical, while 75.6% used a commercial business to perform the acaricide application (Connecticut Department of Public Health, unpublished data). With the phaseout of residential use of the organophosphates chlorpyrifos and diazinon by the US Environment Protection Agency, fewer products will be available to individual homeowners for the control of ticks. With the effectiveness of synthetic pyrethroids, these chemicals will increasingly become the acaricide of choice in area tick control. The synthetic pyrethroid cyfluthrin is particularly effective at a rate of application (0.1 kg a.i. ha–1) 6–45 times less than organophosphate and carbamate materials (Curran et al., 1993; K. Stafford, unpublished data). Solberg et al. (1992) also found spring applications of granular and liquid cyfluthrin (87% and 100% control, respectively, at 2 months) effective against I. scapularis. A single barrier application of granular deltamethrin at woodland edges effectively reduces I. scapularis nymphal abundance in residential settings (Schulze et al., 2001; K. Stafford, unpublished data). However, the application of synthetic chemicals is unacceptable to a substantial proportion of the public in some areas. Non-residual, less toxic chemical control products such as desiccants and insecticidal soaps have been found to significantly reduce nymphal and adult I. scapularis in laboratory and small field trials when containing pyrethrin (Allan and Patrican, 1995; Patrican and Allan, 1995a,b). Nymphs crawling on to landscape stones treated with pyrethrin-based desiccants and insecticidal soaps suffer high (> 88%) mortality (Patrican and Allan, 1995b). While a residential application of natural pyrethrin did not provide effective control of I. scapularis nymphs, K. Stafford (unpublished data) found that a mixture of pyrethrin, the synergist piperonyl butoxide and insecticidal soap was effective in suppressing these ticks. However, more frequent applications of pyrethrin-based products may be required because of a lack of residual activity. Properly timed and targeted applications (i.e. single application as a barrier treatment or application to xeric landscaping substrates in maintained beds and borders near wooded areas) can minimize human exposure, reduce non-target and other potential environmental effects, and yet provide significant levels of tick control.

Host-targeted acaricides The application of acaricides to small mammalian hosts has also been examined for the control of ticks. The relatively small amount of active chemical needed

Environmental Management for LB Control

315

for host-targeted approaches to control LB is an attractive alternative to traditional chemical control because of decreased area-wide pesticide applications and decreased risk of exposure to non-target species. Originally used for the control of plague vectors (Kartman, 1958), Sonenshine and Haines (1985) found that an acaricide bait-box treatment station for the control of immature American dog ticks, Dermacentor variabilis, reduced infestations of rodent populations by > 90%. Frequently, a single species or small complex of species serve as principal hosts for the immature stages or as reservoirs for the bacterial agents of LB. The commercial development of host-targeted permethrin-treated cotton was the first approach for a vector of LB, I. scapularis. Permethrin-treated cottonballs are aimed at immature stages of I. scapularis on white-footed mice, which collect the cotton as nesting material from cardboard tubes distributed approximately every 9–10 m throughout the mouse habitat. Several studies documented significant reductions in tick infestations on whitefooted mice in treated sites (Mather et al., 1987b; Deblinger and Rimmer, 1991; Stafford, 1992b). Mixed results on mice were obtained in studies in New York (Daniels et al., 1991; Ginsberg, 1992). Nymphal reductions on both mice and in host-seeking ticks were also reported in a Massachusetts study with the treatment of one 7.3 ha site (Deblinger and Rimmer, 1991). However, studies in Connecticut and New York State failed to show any reduction in the number of infected, host-seeking I. scapularis nymphs when this product was used for a 3 year period in woodland and residential areas of about 1.6 ha or less (Daniels et al., 1991; Stafford, 1992b). The failure of the tubes of permethrin-treated cotton to significantly affect the tick population in some areas may be due to the presence of alternative nesting materials, alternative tick hosts and/or temporal limitations on the effectiveness of the permethrin. Immigration of untreated tick-infested hosts may also reduce the impact of the permethrin-treated cotton in small, residential applications of the tubes. Mejlon et al. (1995) noted that the application of permethrin-treated cotton reduced numbers of larval I. ricinus on bank voles in Sweden and that significantly fewer nymphs were subsequently collected in the treated areas. However, no impact on larval infestations on the yellow-necked field mouse, Apodemus flavicollis, was observed and the authors noted that, with the diverse I. ricinus host species capable of transmitting B. burgdorferi s.l., the usefulness of rodent-targeted acaricides on nesting material would be limited. Another recent approach, using bait boxes for the topical treatment of rodents with fipronil, is being evaluated for the control of I. scapularis on wild white-footed mice (patents pending, Aventis Environmental Science and Centers for Disease Control and Prevention). Discovered in 1987, fipronil is the first of a new class of insecticides, the phenyl pyrazoles, labelled for a broad range of insect pests on more than 50 crops and various non-crop uses (Aventis, 2000). Fipronil is a reversible inhibitor of the gamma-aminobutyric acid (GABA)-regulated chloride channel that displays selectively tighter binding in insects than in mammals. Percutaneous passage of the compound is low, with the compound concentrating in the epidermis, hair follicle and sebaceous glands (Cochet et al., 1997). Fipronil is commercially available for the control of fleas and ticks on

316

K.C. Stafford and U. Kitron

dogs and cats, with tick control lasting at least a month with a single application. A single topical application of fipronil to a mouse can kill ticks on the animal for up to 42 days (Maupin, 1999), reducing the need for frequent reapplication to the host. Theoretically, a single treatment at a bait box could render a whitefooted mouse virtually tick-free and continue killing any ticks contacting the mouse for over a month. At the time of this writing, trials with a fipronil bait box have been expanded in Connecticut and the northeastern USA and a commercial version of the bait box may be available within a year or two. The acaricide treatment of large wild animal hosts to kill feeding female ticks and thereby control free-living ticks is a relatively recent approach (Pound et al., 1994; Sonenshine et al., 1996). Ivermectin, an avermectin derivative from the soil bacterium Streptomyces avermitilis, is a broad-spectrum systemic parasiticide effective against nematodes, insects, mites and ticks, including A. americanum, B. microplus and I. scapularis. The potential of acaricide treatment of white-tailed deer was demonstrated by the control of A. americanum with ivermectin (Miller et al., 1989; Pound et al., 1996). Although an attempt to control I. scapularis by providing ivermectin-treated maize to free-ranging white-tailed deer on an isolated island did not succeed in reducing free-living tick abundance, > 90% control was obtained on animals with a sufficient serum concentration of ivermectin (Rand et al., 2000). Difficulties in estimating the size of the island’s deer population resulted in insufficient treatment of the host population. The use of a systemic agent in the USA and Europe is unlikely because of the withdrawal period required for clearance of a drug in food animals prior to human consumption. Adult I. scapularis activity and required timing of deer treatments corresponds with the autumn deer-hunting season in the USA. A pesticide self-application device for the control of ticks on white-tailed deer has been developed and patented by scientists at the US Department of Agriculture Livestock Insects Research Laboratory in Kerrville, Texas (US patent no. 5,367,983). The topical pesticide self-application device, termed the 4-poster, consists of a central bin that contains bait to attract deer with one trough and two vertical pesticide impregnated rollers on each end (Pound et al., 1994). Deer contact one of the rollers as they feed, treating the head, neck and ears with an acaricide (Fig. 12.3). Ninety-two and 94% control of nymphal and larval lone star ticks on deer in a large (22.3 ha) deer-fenced pasture in comparison with an untreated pasture was obtained with 2% amitraz (an acaricide with no withdrawal period) applied to deer during the tick season over 3 consecutive years (Pound et al., 2000). The efficacy of the 4-poster and amitraz in controlling I. scapularis feeding on deer and free-living populations of the tick is being evaluated in restricted and unrestricted forested, suburban settings in several states in the northeastern USA. Although adjustments had to be made to ensure sufficient delivery of product to northern deer with their heavy hair coats and larger appetites, and availability of acorns in the autumn at some of the sites affected usage of the device during varying years of the study, generally over 90% of deer in the targeted areas visited the 4-poster (K. Stafford, unpublished data). In Maryland, preliminary analysis has shown that nymphal black-legged

Environmental Management for LB Control

317

Fig. 12.3. (a) The ‘4-poster’ passive topical treatment device for controlling ticks on white-tailed deer with a central bait bin and two feeding and acaricide application stations on either side of the bin. Three of the four vertical applicator rollers are visible. (b) Close-up view of a feeding station showing an adjustable overhanging plate to force deer to rub the sides of the head and neck against the rollers. Only a small amount of bait maize exits from the bait port and deer must place their heads in the required predetermined position to feed (Pound et al., 2000).

ticks were down by 68–71% after the third year of treatment (McGraw and McBride, 2001). Similar results have been observed with lone star ticks in New Jersey (McGraw and McBride, 2001) and with I. scapularis in Connecticut (K. Stafford, unpublished data). The application of acaricides to white-tailed deer is the only approach at this time that can conceivably reduce ticks and have an impact on LB on a community-wide scale, with minimal participation by most residents and with minimal environmental concerns. Whether this approach would be applicable to common European roe deer, Capreolus caprelolus, is unknown and implementation may be more difficult in Europe. The I. ricinus–roe deer system in Europe is less well understood and roe deer are a smaller species, which are broadly dispersed and therefore probably harder to target with a hosttreatment system.

318

K.C. Stafford and U. Kitron

Biological control Ticks have few natural enemies and a number of studies have explored the occurrence and potential use of predators, parasites and pathogens for the control of ticks. There have been very few attempts to control ticks with natural enemies (reviewed by Mwangi et al., 1991). Hu et al. (1998), Samish and Rehacek (1999) and Samish and Alekseev (2001) have reviewed the role of parasitoids, pathogens and predators and arthropod predators on ticks, respectively. Arthropod predators are generally non-specific feeders. Most reports consist of sporadic observations and many observations are based on laboratory observations within a container. Ants, beetles and spiders appear to be the primary arthropods preying on ticks and engorged ticks appear to be a more common target than unfed stages. Samish and Alekseev (2001) reported that 33% of the publications on arthropod predators of ticks were on Ixodes spp., mainly on work from the former USSR, and in contrast with some other genera, the genus Ixodes was preyed upon mainly by beetles rather than ants. Fire ants (Solenopsis geminata) feed on engorged ticks and may have an impact on populations of Boophilus (Barre et al., 1991), and the red imported fire ant, Solenopsis invicta, has severely affected populations of the lone star tick (Harris and Burns, 1972; Burns and Melancon, 1977; Fleetwood et al., 1984). Rodents, lizards, birds and chickens have all been reported to prey on ticks. Milne (1950) noted that predation of preovipositional female I. ricinus was high, with shrews and, to a lesser extent, birds the major predators. The best-known bird predation is probably by the oxpeckers in Africa (Moreau, 1933). Poultry readily eat I. ricinus dropped in sheep pens and chickens will consume attached ticks from cattle and unfed ticks from the vegetation (Milne, 1950; Hassan et al., 1991). Anecdotal reports suggest that chickens and guineafowl may have an impact on local abundance of I. scapularis, but it remains unclear what effect these predators have on the tick population. Ticks are parasitized by seven known species of minute (2 mm) chalcid wasps in the family Encyrtidae, which have been recovered from many species of ticks worldwide (Davis, 1986; Hu et al., 1998). Two species of these insects were first recognized and described in the USA, Ixodiphagus hookeri (Howard) and Ixodiphagus texanus Howard. The first species was mass-released in the late 1920s and 1930s on the Elizabeth Islands and Martha’s Vineyard, Massachusetts, to control the American dog tick, D. variabilis, and in Montana and other western states to control the Rocky Mountain wood tick, Dermacentor andersoni, a vector of Rocky Mountain spotted fever (Larrouse et al., 1928; Cooley and Kohls, 1934; Smith and Cole, 1943). No reduction in tick abundance due to the parasitoids was observed and few wasps were recovered following these releases, although the wasp apparently became established in I. scapularis on Naushon Island (Larrouse et al., 1928; Cobb, 1942; Mather et al., 1987a). The failures of the previous releases decades ago may be attributed to some ecological constraint on the parasitoid, timing of the release, climatic incompatibility, or limitation on parasitoid reproduction in the targeted tick species. The report by Mather et al. (1987a) that the prevalence of B. burgdorferi or

Environmental Management for LB Control

319

B. microti, the cause of human babesiosis, was nearly 40% lower in host-seeking nymphs of I. scapularis parasitized by I. hookeri renewed interest in these parasitoids in vector control, and more recently a theoretical case has been made for the control of I. scapularis by parasitoid augmentation, with a call for a major research initiative (Knipling and Steelman, 2000). Mwangi et al. (1997) reported that nymphal Amblyomma variegatum declined by 95% following the monthly release over a year of 150,000 I. hookeri on ten head of cattle in a 4 ha pasture. A parasitism rate of 51% in the nymphs was obtained with the high parasitoid release rate, but few (9%) nymphs were parasitized during the post-release period of the study. These wasps are established at several insular locations in New England and New York where deer and tick populations are extremely high (Hu et al., 1993; Stafford et al., 1996), suggesting that the parasitoid is strongly dependent on densities of ticks or tick hosts. The reduction of deer at two insular populations of the black-legged tick in Connecticut has resulted in dramatic declines in the presence of I. hookeri (K. Stafford, unpublished data), suggesting that the usefulness of this parasitoid to control I. scapularis may be limited in any integrated control programme. The application of entomopathogenic fungi to tick habitat or vertebrate hosts to control ticks is another environmentally acceptable approach, and fungi are one of the more promising biological control agents for ticks. A number of fungi have been isolated from ticks, primarily engorged females from deer, although some fungal species have been found on larvae, nymphs and unfed females. These include Paecilomyces farinosus, Verticillium lecanii, Beauveria bassiana and some other species from these three genera in the Hyphomycetes family. Recovery rates are low, however, generally less than 2–5%. V. lecanii appears to be a common fungal pathogen of I. scapularis and I. ricinus in nature (Kalsbeek et al., 1995; Zhioua et al., 1999). Entomopathogenic fungi such as B. bassiana and Metarhizium anisopliae have been reported to cause mortality in several tick species such as Rhipicephalus appendiculatus, A. variegatum (Kaaya et al., 1996), I. ricinus (Samsinakova et al., 1974) and I. scapularis (Zhioua et al., 1997). The role of these fungi in the control of natural populations is unknown, but appears low. Even though local strains isolated from ticks may be more pathogenic, existing commercial formulations of entomopathogenic fungi might be used to control ticks. Two products, BotaniGard (Mycotech, Butte, Montana) and Naturalis T & O (Troy Biosciences, Phoenix, Arizona), are labelled in the USA for ornamental and turf applications. Perimeter treatments in trials at residential sites with these two strains of B. bassiana reduced nymphal I. scapularis by 59–89% in one trial and 80–90% of a sample of nymphs developed mucoses (K. Stafford and S. Allan, unpublished data). Although ticks do not support a full nematode life cycle, nematode parasites of insects have been proposed as possible biological control agents for ticks (Samish and Glazer, 1991; Samish and Rehacek, 1999). The engorged females of several tick species have been found to be susceptible to infection with entomopathogenic steinernematid and heterorhabiditid nematodes (Samish et al., 1999, 2000). Juvenile nematodes penetrate the host by the genital pore in female

320

K.C. Stafford and U. Kitron

ticks, and release symbiotic bacteria, which kill the host. Susceptibility to nematodes varies with the species of tick, developmental stage of the tick and nematode strain. Zhioua et al. (1995) and Hill (1998) demonstrated that engorged I. scapularis females, but not unfed females, engorged nymphs or engorged larvae, are susceptible to steinernematid nematodes. The application of nematodes to engorged female ticks will kill these ticks prior to laying eggs and could potentially reduce tick reproduction. However, sensitivity to autumn temperatures and lack of susceptibility of unfed female ticks limit the potential application of nematodes for the vectors of LB at present. Microsporidian parasites (Nosema spp.) have been recovered from I. ricinus and, although microsporidia have potential as biocontrol agents for a number of pests, their role in ticks remains unknown (Weiser and Rehacek, 1975).

Pheromones Tick pheromones and related chemicals are potential tools that could be applied to tick management by disrupting tick behaviour and reproduction. Pheromones are messenger chemicals used by animals to modify behaviour in members of the same species. A messenger chemical that benefits only the recipient, such as a host odour attractive to ticks or to parasitoids, is referred to as a kairomone. Pheromones used by various tick species to control and modulate tick movement and mating behaviour are broadly classified as assembly, aggregation-attachment or sex-attractant pheromones. For a more detailed review of pheromone types and methods for the identification of tick pheromones, readers are referred to Sonenshine (1991). Assembly pheromones are present in many, although not all, tick species and induce clustering of free-living ticks in their environment. Aggregationattachment pheromones are produced by males of several Amblyomma spp. and strongly attract unfed adult, sometimes nymphal, ticks. They induce attachment in areas where male ticks are feeding, and permit identification of infested hosts by questing ticks, clustering of ticks on the host and facilitation of mating (Norval et al., 1989). The mating process in ticks is governed by several sex pheromones in a discrete series of phases. The hierarchy of the mating sequence in ticks other than the genus Ixodes is relatively consistent as egg or sperm development proceeds only following attachment and commencement of feeding (for a full review of tick reproduction, see Sonenshine, 1994). In Ixodes spp., mating may occur off the host in the vegetation or on the host, because male gamete development is not dependent on tick feeding. Little is known about pheromonal regulation of behaviour in Ixodes spp. ticks. The chemical 2,6-dichlorophenol is the sex attractant secreted by at least 14 species of ticks (Hamilton, 1992), but is not the sex attractant for Ixodes spp. ticks, which still need to be characterized. The existence of assembly and sex pheromones has been reported for I. ricinus (Graf, 1975; Hajkova and Leahy, 1982). The presence of assembly pheromones, which appear to be purines or purine-like substances, attractive to nymphal and adult

Environmental Management for LB Control

321

I. scapularis, has recently been documented using bioassays of cast skins and exudates from moulted ticks (Allan, 1999; Allan and Sonenshine, 2002). Carroll (2001; Carroll et al., 1996) has reported that adult I. scapularis exhibit an arrestant response to host-derived chemicals (e.g. tarsal and interdigital gland) secreted by white-tailed deer in the laboratory and the field. While the distance at which ticks can detect and respond to these substances is unknown, kairomones could serve to concentrate ticks in areas with increased likelihood of finding a host. Research has focused on use of pheromones as a lure to attract ticks to a toxicant, to disrupt mating behaviour, or in a baited decoy system to attract and kill male ticks (Sonenshine, 1994). A combination of a pheromone and an acaricide could kill ticks more effectively and selectively because ticks are attracted by the pheromone to the acaricide. The combination of 2,6-dichlorophenol with an acaricide kills mate-seeking male D. variabilis and disrupts the mating process (Sonenshine and Haines, 1985; Sonenshine et al., 1992). An attractant decoy tag containing an acaricide–pheromone mixture on collars and/or the tails of cattle has been shown to provide excellent control of bont ticks (Amblyomma hebraeum and A. variegatum) for many weeks (Norval et al., 1996; Allan et al., 1998). A controlled-release formulation containing an assembly pheromone or a kairomone and an acaricide could potentially be used against free-living ticks and improve area control of LB vectors with better efficacy and less toxicity than current applications.

Integrated tick management There are many definitions of IPM, but it basically involves the selection, integration and implementation of several pest control actions based on predicted ecological, economic and sociological consequences (Axtell, 1981) with the goal of reduction, rather than elimination, of the pest population. Two of the basic concepts of IPM are the economic injury level, the pest population density at which economic losses exceed the cost of control, and the economic threshold, the pest density at which control measures should be applied to prevent the pest population from reaching the injury level. From a public health perspective, this implies an acceptable density of vector ticks and risk of LB before management practices are implemented. In the case of LB, eradication of the tick vector is not feasible and acceptable thresholds would be different for residential areas, recreational areas and less visited parts of woodland parks, and tied to the perception of risk as well as the actual density of ticks required for transmission to humans. In terms of human nuisance, for example, an economic threshold of 0.65 ticks h–1 in a carbon dioxide sample was proposed for the lone star tick based on an attack rate of less than one tick per human visitor per day (Mount and Dunn, 1983). It is difficult to put a value on a human disease like LB and, consequently to define an acceptable ‘economic’ level for managing the disease. Nevertheless, individuals make this decision when they elect to implement or obtain tick control services, alter their landscape or seek vaccination. As noted

322

K.C. Stafford and U. Kitron

earlier, the use of personal protection measures is influenced by both the perception of risk and the degree of inconvenience involved in the protective activity. The basic aim of any LB prevention programme is to reduce the number of human cases within a community or larger geographical region. On a practical basis, reducing the risk of tick bite, and consequently disease, has been limited to personal protection measures and the residential use of acaricides. The use of many technologies (e.g. host immunity, anti-tick vaccines, pheromones, parasites and pathogens) is being researched or remains speculative and others (e.g. acaricide on mouse nesting material) have had mixed or negative results (Daniels et al., 1991; Ginsberg, 1992; Stafford, 1992b). The combination of methods offers the potential for the greatest level of tick management. For example, results of a project to manage lone star tick populations in a recreation area demonstrated that combinations of three tick control methods – acaricides, vegetation management and exclusion of white-tailed deer by a fence – provided greater tick reductions than each method alone (Bloemer et al., 1990). All three methods together produced the greatest reduction, with 92–99% control of lone star larvae, nymphs and adults. Computer models of the life cycle of a number of tick species (Haile and Mount, 1987; Mount et al., 1997a) have been used to study various combinations of tick management strategies, which could assist in the development of tick management programmes (Haile et al., 1990; Mount et al., 1997b). The LYMESIM model has been extended to the simulation of several management strategies on I. scapularis densities (Mount et al., 1997a,b). Area-wide acaricide, vegetation reduction or a combination of the two was the most effective in small recreational or residential sites. Simulated 50 and 100% reductions in vegetation were highly effective in reducing infected nymphal abundance (89 and > 99% by year 3, respectively). Simulations of area-wide acaricides applied annually to each stage reduced larvae, nymphs and adults by 82, 91 and 93%, respectively. The combination of acaricides and vegetation reduction in the LYMESIM simulation produced a 96% reduction in the seasonal abundance of infected nymphal I. scapularis, comparable with that observed with A. amercainum nymphs in experiments in Kentucky, Tennessee and Oklahoma (Mount and Whitney, 1984; Bloemer et al., 1990). The acaricidal treatment of deer was the most cost-effective, long-term strategy for larger areas, but the model also suggested deer density reductions to be an important element. The simulations suggested that it might take 4–10 years to reduce the population of I. scapularis below the threshold necessary to maintain spirochaetal transmission (Mount et al., 1997b). LYMESIM was also used to assess the effectiveness of eight interventions in a hypothetical community of 10,000 people over 4 years with an annual incidence rate of 0.01 cases per person per year, i.e. 1000/100,000 population (Hayes et al., 1999), a high rate that has been noted in some Connecticut communities (Stafford et al., 1998; Connecticut Department of Public Health, 2001). The interventions simulated were area-wide acaricide, acaricide to nesting mice,

Environmental Management for LB Control

323

acaricide to white-tailed deer, removal of white-tailed deer, fencing out deer, reduction of habitat vegetation, repellent and tick checks and the Lyme vaccine. Various estimates of intervention use were based, in part, on published surveys on LB-related behaviours of people and experiment efficacy of the intervention. The number of LB cases prevented was determined under varying assumptions for a baseline, worst-case and best-case scenario for each intervention. The impact of the interventions was dependent upon the level of engagement. For example, if only 50% of the population is vaccinated, 74 cases of LB would be prevented, but with 95% vaccinated, 361 cases would be prevented. Reduction in tick habitat vegetation by 40% of residents with 90% habitat reduction on lawns, 80% habitat reduction in ecotone and 10% reduction in forested areas resulted in the prevention of only 94 cases. Acaricidal treatment of nesting mice prevented the fewest cases, while the acaricidal treatment of deer prevented the most cases (113–306) of LB under all scenarios except the best use of the Lyme vaccine. Use of an LB vaccine only protects vaccinated individuals and is dependent on a high level of engagement, while some other interventions may protect neighbours who do not engage in these interventions. Removal or exclusion of deer also decreases disease incidence, although risk increased for the first 2 years, reflecting an increase in the proportion of infected nymphs.

The Future for Management of Lyme Borreliosis An integrated approach to the control of LB is at an early stage of development, with the objective of lowering tick densities and reducing the proportion of ticks infected with B. burgdorferi s.l. The choice of intervention strategies used in an integrated programme within a disease-endemic community will depend on the level of engagement by that community. Not all methods may be applicable or acceptable to a specific target population. Some interventions only require the participation of individual households while others will require the cooperation of the entire community. Interventions such as the vaccine or personal protection measures will only reduce risk for the individual practising such behaviours. The application of acaricides to the environment or tick hosts may benefit neighbouring properties and visitors to the treated properties. Current technologies, properly applied, clearly could reduce the incidence of LB. Nevertheless, utilization of existing technology, particularly acaricides, is low and novel strategies for the control of ticks and the pathogens they transmit are needed. Existing technologies need to be more intensively applied in areas endemic for LB. Geographical areas of high risk for LB can be identified by human case reports, records of tick bite and mapping of natural foci for the pathogen–vector–host cycle based on a sound understanding of the epidemiology and ecology of the disease and known human activity patterns. Remote sensing and GIS technologies can be used to identify and map the risk of disease and target intervention efforts. Satellite imagery, GIS and spatial statistics are tools that can contribute more to the design and implementation of LB management

324

K.C. Stafford and U. Kitron

programmes. For example, a detailed geo-referenced database can be used to determine which control strategies will be applied, and to what extent, in different parts of a control area. As long as deer and other hosts remain abundant, tick vectors increase in abundance or spread geographically and more people either live by or visit tick-infested areas, the incidence of LB and other tick-associated diseases will increase. LB is, or could be, a largely preventable disease. Several experimental community or population-based prevention projects have recently been initiated in the northeastern USA, but LB prevention studies or tick control programmes in Europe or Russia are few or non-existent. In the absence of an effective LB programme, wider acceptance of the current LB vaccine (in the USA), or development of more effective broadly applicable vaccines or better tick management techniques, LB will continue to be an emerging disease problem. The application of tick-management programmes for the control of LB will ultimately be a policy choice based on the risk or perceived risk of LB and the cost and effectiveness of an intervention programme. The environmental management of LB will require government-based programmes supported by public funds, similar to mosquito and mosquito-borne disease programmes that have existed in the USA for decades, and will integrate education, vaccine and various strategies for tick control. Public health initiatives to reduce LB in selected, targeted communities in the USA are an important step in that direction. At the current time, the reduction of LB and other tick-associated diseases will require a multifaceted approach utilizing a variety of measures for tick management and disease prevention by both the individual and the community.

References Adler, G.H., Telford, S.R. III, Wilson, M.L. and Spielman, A. (1992) Vegetation structure influences the burden of immature Ixodes dammini on its main host, Peromyscus leucopus. Parasitology 105, 105–110. Alekseev, A.N., Burenkova, L.A., Vasilieva, I.S., Dubinina, H.V. and Chunikhin, S.P. (1996) Preliminary studies on virus and spirochete accumulation in the cement plug of ixodid ticks. Experimental and Applied Acarology 20, 713–723. Allan, S.A. (1999) Pheromonal attraction in ticks. In: Proceedings of the National Conference on the Prevention and Control of Tick-borne Disease, 7–9 March, 1999, New York. American Lyme Disease Foundation, p. 18.

Allan, S.A. and Patrican, L.A. (1995) Reduction of immature Ixodes scapularis (Acari: Ixodidae) in woods by application of desiccant and insecticidal soap formulations. Journal of Medical Entomology 32, 16–20. Allan, S.A. and Sonenshine, D.E. (2002) Evidence of an assembly pheromone in the black-legged deer tick, Ixodes scapularis. Journal of Chemical Ecology 28, 15–27. Allan, S.A., Barre, N., Sonenshine, D.E. and Burridge, M.J. (1998) Efficacy of tags impregnated with pheromone and acaricide for the control of Amblyomma variegatum. Medical Veterinary Entomology 12, 141–150. Anderson, J.F., Johnson, R.C., Magnarelli,

Environmental Management for LB Control

L.A., Hyde, F.W. and Myers, J.E. (1987) Prevalence of Borrelia burgdorferi and Babesia microti in mice on islands inhabited by white-tailed deer. Applied and Environmental Microbiology 53, 892–894. Aventis (2000) Fipronil Worldwide Technical Bulletin. Aventis Crop Science, Lyon, France, 27 pp. Axtell, R.C. (1981) Livestock Integrated Pest Management (IPM): principals and prospects. In: Knapp, F.W. (ed.) Systems Approach to Animal Health and Production: a Symposium. University of Kentucky, pp. 31–40. Barnard, D.R. (1986) Density perturbation in populations of Amblyomma americanum (Acari: Ixodidae) in beef cattle forage areas in response to two regimens of vegetation management. Journal of Economic Entomology 79, 122–127. Barre, N., Maulson, H., Harris, G.I. and Kermarrec, A. (1991) Predators of the tick Amblyomma variegatum (Acari: Ixodidae) in Guadeloupe, French West Indies. Experimental and Applied Acarology 12, 163–170. Bloemer, S.R., Snoddy, E.L., Cooney, J.C. and Fairbanks, K. (1986) Influence of deer exclusion on populations on lone star ticks and American dog ticks (Acari: Ixodidae). Journal of Economic Entomology 79, 679–683. Bloemer, S.R., Mount, G.A., Morris, T.A., Zimmerman, R.H., Barnard, D.R. and Snoddy, E.L. (1990) Management of lone star ticks (Acari: Ixodidae) in recreational areas with acaricide applications, vegetative management, and exclusion of white-tailed deer. Journal of Medical Entomology 27, 543–550. Bowman, A.S., Coons, L.B., Needham, G.R. and Sauer, J.R. (1997) Tick saliva: recent advances and implications for vector competence. Medical and Veterinary Entomology 11, 277–285. Brown, S.W., Cartter, M.L., Hadler, J.L. and Hooper, P.F. (1992) Lyme disease knowledge, attitudes, and behaviors –

325

Connecticut, 1992. Morbidity and Mortality Weekly Report 41, 505–507. Burns, E.C. and Melancon, D.G. (1977) Effect of imported fire ant (Hymenoptera: Formicidae) invasion on lone star tick (Acari: Ixodidae) populations. Journal of Medical Entomology 14, 247–249. Carroll, J.F. (2001) Interdigital gland substances of white-tailed deer and the response of host-seeking ticks (Acari: Ixodidae). Journal of Medical Entomology 38, 114–117. Carroll, J.F., Mills, G.D. Jr. and Schmidtmann, E.T. (1996) Field and laboratory responses of adult Ixodes scapularis (Acari: Ixodidae) to kairomones produced by white-tailed deer. Journal of Medical Entomology 33, 640–644. Cartter, M.L., Farley, T.A., Ardito, H.A. and Hadler, J.L. (1989) Lyme disease prevention – knowledge, beliefs, and behaviors among high school students in an endemic area. Connecticut Medicine 53, 354–356. Centers for Disease Control and Prevention (1999a) Notice to readers: availability of Lyme disease vaccine. Morbidity and Mortality Weekly Report 48, 35–36, 43. Centers for Disease Control and Prevention (1999b) Recommendations for the use of Lyme disease vaccine: recommendations of the Advisory Committee on Immunization Practices (ACIP). Morbidity and Mortality Weekly Report 48 (Suppl. RR-7), 1–26. Cobb, S. (1942) Tick parasites on Cape Cod. Science 95, 503. Cochet, P., Birckel, P., Bromet-Petit, M., Bromet, N. and Weil, A. (1997) Skin distribution of fipronil by microautoradiography following topical administration to the beagle dog. European Journal of Drug Metabolism and Pharmacokinetics 22, 211–216. Connecticut Department of Public Health (2001) Lyme Disease Statistics, 1996–2000. www.state.ct.us/dph, 4.

326

K.C. Stafford and U. Kitron

Cooley, R.A. and Kohls, G.M. (1934) A summary of tick parasites. 5th Pacific Science Congress Proceedings 5, 3375–3381. Cortinas, M.R., Guerra, M.A., Jones, C.J. and Kitron, U. (2002) Characterization and prediction of tick-borne disease foci. International Journal of Medical Microbiology 291(33) (Suppl.). Costello, C.M., Steere, A.C., Pinkerton, R.E. and Feder, H.M. (1989) A prospective study of tick bites in an endemic area for Lyme disease. Journal of Infectious Diseases 159, 136–139. Curran, K.L., Fish, D. and Piesman, J. (1993) Reduction of nymphal Ixodes dammini (Acari: Ixodidae) in a residential suburban landscape by area application of insecticides. Journal of Medical Entomology 30, 107–113. Daniel, M. and Kolar, J. (1990) Using satellite data to forecast the occurrence of the common tick Ixodes ricinus (L.). Journal of Hygiene, Epidemiology, Microbiology and Immunology 34, 243–252. Daniel, M., Kolar, J., Zeman, P., Pavelka, K. and Sadlo, J. (1998) Predictive map of Ixodes ricinus high-incidence habitats and a tick-borne encephalitis risk assessment using satellite data. Experimental and Applied Acarology 22, 417–433. Daniels, T.J. and Fish, D. (1995) Effect of deer exclusion on the abundance of immmature Ixodes scapularis (Acari: Ixodidae) parasitizing small and medium-sized mammals. Journal of Medical Entomology 32, 5–11. Daniels, T.J., Fish, D. and Falco, R.C. (1991) Evaluation of host-targeted acaricide for reducing risk of Lyme disease in southern New York state. Journal of Medical Entomology 28, 537–543. Daniels, T.J., Fish, D. and Schwartz, I. (1993) Reduced abundance of Ixodes scapularis (Acari: Ixodidae) and Lyme disease risk by deer exclusion. Journal of Medical Entomology 30, 1043–1049. Davis, A.J. (1986) Bibilography of the Ixodiphagini (Hymenoptera, Chalcidoidea, Encyrtidae), parasites of ticks

(Acari, Ixodidae), with notes on their biology. Tijdschrift voor Entomologie 129, 181–190. Deblinger, R.D. and Rimmer, D.W. (1991) Efficacy of permethrin-based acaricide to reduce the abundance of Ixodes dammini (Acari: Ixodidae). Journal of Medical Entomology 28, 708–711. Deblinger, R.D., Wilson, M.L., Rimmer, D.W. and Spielman, A. (1993) Reduced abundance of immature Ixodes dammini (Acari: Ixodidae) following incremental removal of deer. Journal of Medical Entomology 30, 144–150. DeNicola, A.J., VerCauteren, K.C., Curtis, P.D. and Hygnstrom, S.E. (2000) Managing White-tailed Deer in Suburban Environments. Cornell Cooperative Extension, Ithaca, New York, 52 pp. Dieter, S.W., Fish, D., Bros, S.M., Frank, D.H. and Wood, B.L. (1997) Landscape characterization of peridomestic risk for Lyme disease using satellite imagery. American Journal of Tropical Medicine and Hygiene 57, 687–692. Ellingwood, M.R. and Caturano, S.L. (1988) An Evaluation of Deer Management Options. Publication no. DR-11, Connecticut Department of Environmental Protection, Wildlife Division, Hartford, Connecticut, 12 pp. Estrada-Peña, A. (1998) Geostatistics and remote sensing as predictive tools of tick distribution: a cokriging system to estimate Ixodes scapularis (Acari: Ixodidae) habitat suitability in the United States and Canada from advanced very high resolution radiometer satellite imagery. Journal of Medical Entomology 35, 989–995. Estrada-Peña, A. (2001) Forecasting habitat suitability for ticks and prevention of tick-borne diseases. Veterinary Parasitology 98, 111–132. Fish, D. and Howard, C.A. (1999) Methods used for creating a national Lyme disease risk map. Morbidity and Mortality Weekly Report 48, 21–24. Fix, A.D., Strickland, G.T. and Grant, J.

Environmental Management for LB Control

(1998) Tick bites and Lyme disease in an endemic setting: problematic use of serologic testing and prophylactic antibiotic therapy. Journal of the American Medical Association 279, 206–210. Fleetwood, S.C., Teel, P.D. and Thompson, G. (1984) Impact of imported fire ant on lone star tick mortality in open and canopied pastured habitats of east central Texas. Southwest Entomologist 9, 158–162. Fricks, L.D. (1915) Rocky Mountain spotted fever. Public Health Reports 30, 148–165. Gern, L., Lebet, N. and Moret, J. (1996) Dynamics of Borrelia burgdorferi infection in nymphal Ixodes ricinus ticks during feeding. Experimental and Applied Acarology 20, 649–658. Gern, L., Hu, C.M., Voet, P., Hauser, P. and Lobet, Y. (1997) Immunization with a polyvalent OspA vaccine protects mice against Ixodes ricinus tick bites infected by Borrelia burgdorferi ss, Borrelia garinii, and Borrelia afzelii. Vaccine 15, 1551–1557. Ginsberg, H.S. (1992) Ecology and Management of Ticks and Lyme Disease at Fire Island National Seashore and Selected National Parks. Science Monograph NPS/NRSUNJ/NRSM-92/20, US Department of the Interior, National Park Service, 77 pp. Ginsberg, H.S. and Zhioua, E. (1999) Influence of deer abundance on the abundance of questing adult Ixodes scapularis (Acari: Ixodidae). Journal of Medical Entomology 36, 379–381. Glass, G.E., Amerasinghe, F.P., Morgan, J.M. III and Scott, T.W. (1994) Predicting Ixodes scapularis abundance on white-tailed deer using geographic information systems. American Journal of Tropical Medicine and Hygiene 51, 538–544. Glass, G.E., Schwartz, B.S., Morgan, J. M. III, Johnson, D.T., Noy, P.M. and Israel, E. (1995) Environmental risk factors for Lyme disease identified with

327

geographic information systems. American Journal of Public Health 85, 944–948. Graf, J.F. (1975) Ecologie and ethologie d’Ixodes ricinus L. en Suisse (Ixodoidea: Ixodidae). Cinquième note: mise en évidence d’une phéromone sexuelle chez Ixodes ricinus. Acarologia 17, 436–441. Gray, J.S., Kahl, O., Janetzki, C. and Stein, J. (1992) Studies on the ecology of Lyme disease in a deer forest in County Galway, Ireland. Journal of Medical Entomology 29, 915–920. Gray, J.S., Kahl, O., Robertson, J.N., Daniel, M., Estrada-Peña, A., Gettinby, G., Jaenson, T.G.T., Jensen, P., Jongejan, F., Korenburg, E., Kurtenbach, K. and Zeman, P. (1998) Lyme borreliosis habitat assessment. Zentralblatt für Bakteriologie 287, 211–228. Guerra, M., Walker, E., Jones, C., Paskewitz, S., Cortinas, M., Stancil, A., Beck, L., Bobo, M. and Kitron, U. (2001a) Predicting the risk of Lyme disease: habitat suitability for Ixodes scapularis in the north-central U.S. Emerging Infectious Diseases 8, 289–297. Guerra, M.A., Walker, E.D. and Kitron, U. (2001b) Canine surveillance system for Lyme borreliosis in Wisconsin and northern Illinois: geographic distribution and risk factor analysis. American Journal of Tropical Medicine and Hygiene 65, 546–592. Gylfe, A., Olsen, B., Strasevicius, D., Ras, N.M., Weihe, P., Noppa, L., Ostberg, Y., Baranton, G. and Bergstrom, S.J. (1999) Isolation of Lyme disease Borrelia from puffins (Fratercula arctica) and seabird ticks (Ixodes uriae) on the Faeroe Islands. Journal of Clinical Microbiology 37, 890–896. Haile, D.G. and Mount, G.A. (1987) Computer simulation of population dynamics of the lone star tick, Amblyomma americanum (Acari: Ixodidae). Journal of Medical Entomology 24, 356–369. Haile, D.G., Mount, G.A. and Cooksey,

328

K.C. Stafford and U. Kitron

L.M. (1990) Computer simulation of management strategies for American dog ticks (Acari: Ixodidae) and Rocky Mountain spotted fever. Journal of Medical Entomology 27, 686–696. Hajkova, Z. and Leahy, M.G. (1982) Pheromone-regulated aggregation in larvae, nymphs, and adults of Ixodes ricinus (L.) (Acarina: Ixodidae). Folia Parasitologica 29, 61–67. Hamilton, J.G.C. (1992) The role of pheromones in tick biology. Parasitology Today 8, 130–133. Harris, W.G. and Burns, E.C. (1972) Predation on the lone star tick by the imported fire ant. Environmental Entomology 1, 362–365. Hassan, S.M., Dipeolu, O.O., Amoo, A.O. and Odhiambo, T.R. (1991) Predation of livestock ticks by chickens. Veterinary Parasitology 38, 199–204. Hayes, E.B., Maupin, G.O., Mount, G.A. and Piesman, J. (1999) Assessing the prevention effectiveness of local Lyme disease control. Journal of Public Health Management and Practice 5, 84–92. Hill, D.E. (1998) Entomopathogenic nematodes as control agents of developmental stages of the black-legged tick, Ixodes scapularis. Journal of Parasitology 84, 1124–1127. Hoch, A.L., Semtner, P.J., Baker, R.W. and Hair, J.A. (1972) Preliminary observation on controlled burning for lone star tick (Acarina: Ixodidae) control in woodlots. Journal of Medical Entomology 9, 446–451. Hu, R., Hyland, K.E. and Mather, T.N. (1993) Occurence and distribution in Rhode Island of Hunterellus hookeri (Hymenoptera: Encyrtidae), a wasp parasitoid of Ixodes dammini. Journal of Medical Entomology 30, 277–280. Hu, R., Hyland, K.E. and Oliver, J.H. (1998) A review on the use of Ixodiphagus wasps (Hymenoptera: Encyrtidae) as natural enemies for the control of ticks (Acari: Ixodidae). Systematic and Applied Acarology 3, 19–28.

Jaenson, T.G.T., Fish, D., Ginsberg, H.S., Gray, J.S., Mather, T.N. and Piesman, J. (1991) Workshop summary: methods for control of tick vectors of Lyme borreliosis. Scandinavian Journal of Infectious Disease Suppl. 77, 151–157. Kaaya, G.P., Mwangi, E.N. and Ouna, E.A. (1996) Prospects for biological control of livestock ticks, Rhipicephalus appendiculatus and Amblyomma variegatum, using the entomogenous fungi Beauveria bassiana and Metarhizium anisopliae. Journal of Invertebrate Pathology 67, 15–20. Kahl, O. and Knülle, W. (1988) Water vapour uptake from subsaturated atmospheres by engorged immature ixodid ticks. Experimental and Applied Acarology 4, 73–83. Kahl, O., Janetzki-Mittmann, C., Gray, J.S., Jonas, R., Stein, J. and de Boer, R. (1998) Risk of infection with Borrelia burgdorferi sensu lato for a host in relation to the duration of nymphal Ixodes ricinus feeding and the method of tick removal. Zentralblatt für Bakteriologie 287, 41–52. Kalsbeek, V., Frandsen, F. and Steenberg, T. (1995) Entomopathogenic fungi associated with Ixodes ricinus ticks. Experimental and Applied Acarology 19, 45–51. Kartman, L. (1958) An insecticide bait-box method for the control of sylvatic plague vectors. Journal of Hygiene 56, 455–465. Kilpatrick, H.J., Spohr, S.M. and Chasko, G.G. (1997) A controlled deer hunt on a state-owned coastal reserve in Connecticut: controversies, strategies, and results. Wildlife Society Bulletin 25, 451–456. Kitron, U. (1998) Landscape ecology and epidemiology of vector-borne diseases: tools for spatial analysis. Journal of Medical Entomology 35, 435–445. Kitron, U. (2000) Risk maps: transmission and burden of vector-borne diseases. Parasitology Today 16, 324–325. Kitron, U. and Kazmierczak, J. (1997) Spatial analysis of the distribution of

Environmental Management for LB Control

Lyme disease in Wisconsin. American Journal of Epidemiology 145, 558–566. Kitron, U., Jones, C.J., Bouseman, J.K., Nelson, J.A. and Baumgartner, D.L. (1992) Spatial analysis of the distribution of Ixodes dammini (Acari: Ixodidae) on white-tailed deer in Ogle County, Illinois. Journal of Medical Entomology 29, 259–266. Knipling, E.F. and Steelman, C.D. (2000) Feasibility of controlling Ixodes scapularis ticks (Acari: Ixodidae), the vector of Lyme disease, by parasitoid augmentation. Journal of Medical Entomology 37, 645–652. Korenberg, E.I. (1994) Comparative biology and epidemiology of Lyme disease and tick-borne encephalitis in the former Soviet Union. Parasitology Today 10, 157–160. Larrouse, F., King, A.G. and Wolbach, S.B. (1928) The overwintering in Massachusetts of Ixodiphagus caucurteri. Science 67, 351–353. Lees, A.D. (1946) The water balance in Ixodes ricinus L. and certain other species of ticks. Parasitology 37, 1–20. Lees, A.D. and Milne, A. (1951) The seasonal and diurnal activities of individual sheep ticks (Ixodes ricinus L.). Parasitology 41, 189–208. Maiwald, M., Oehme, R., March, O., Petney, T.N., Kimmig, P., Naser, K., Zappe, H.A., Hassler, D. and Doeberitz, M.V.K. (1998) Transmission risk of Borrelia burgdorferi sensu lato from Ixodes ricinus ticks to humans in southwest Germany. Epidemiological Infections 121, 103–108. Mather, T.N., Piesman, J. and Spielman, A. (1987a) Absence of spirochetes (Borrelia burgdorferi) and piroplasms (Babesia microti) in deer ticks (Ixodes dammini) parasitized by chalcid wasps (Hunterellus hookeri). Medical and Veterinary Entomology 1, 3–8. Mather, T.N., Ribeiro, J.M.C. and Spielman, A. (1987b) Lyme disease and babesiosis: acaricide focused on poten-

329

tially infected ticks. American Journal of Tropical Medicine and Hygiene 36, 609–614. Mather, T.N., Duffy, D.C. and Campbell, S.R. (1993) An unexpected result from burning vegetation to reduce Lyme disease transmission risks. Journal of Medical Entomology 30, 642–645. Maupin, G.O. (1999) Innovations in acaricides and delivery methods against ticks. In: Proceedings of the National Conference on the Prevention and Control of Tick-borne Disease, 7–9 March, 1999, New York. American Lyme Disease Foundation, p. 16. Maupin, G.O., Fish, D., Zultowsky, J., Campos, E.G. and Piesman, J. (1991) Landscape ecology of Lyme disease in a residential area of Westchester County, New York. American Journal of Epidemiology 133, 1105–1113. McGraw, L. and McBride, J. (2001) Out of the Lyme-light. Agricultural Research 49, 4–7. Mejlon, H.A., Jaenson, T.G.T. and Mather, T.N. (1995) Evaluation of hosttargeted applications of permethrin for control of Borrelia-infected Ixodes ricinus (Acari: Ixodidae). Medical Veterinary Entomology 9, 207–210. Meltzer, M.I., Dennis, D.T. and Orloski, K.A. (1999) The cost effectiveness of vaccinating against Lyme disease. Emerging Infectious Diseases 5, 321–328. Merler, S., Furlanello, C., Chemini, C. and Nicolini, G. (1996) Classification tree methods for analysis of mesoscale distribution of Ixodes ricinus (Acari: Ixodidae) in Trentino, Italian alps. Journal of Medical Entomology 33, 888–893. Miller, J.A., Garris, G.I., George, J.E. and Oehler, D.D. (1989) Control of lone star ticks (Acari: Ixodidae) on Spanish goats and white-tailed deer with orally administered ivermectin. Journal of Economic Entomology 82, 1650–1656. Milne, A. (1950) The ecology of the sheep tick, Ixodes ricinus L.: microhabitat econ-

330

K.C. Stafford and U. Kitron

omy of the adult tick. Parasitology 40, 14–34. Monsen, S.E., Bronson, L.R., Tucker, J.R. and Smith, C.R. (1999) Experimental and field evaluations of two acaracides for control of I. pacificus (Acari: Ixodidae) in northern California. Journal of Medical Entomology 36, 660–665. Moreau, R.E. (1933) The food of the redbilled oxpecker, Buphagus erythrorhynchus (Stanley). Bulletin of Entomological Research 24, 325–335. Mount, G.A. (1981) Control of the lone star tick in Oklahoma parks through vegetative management. Journal of Economic Entomology 74, 173–175. Mount, G.A. and Dunn, J.E. (1983) Economic thresholds for lone star ticks (Acari: Ixodidae) in recreational areas based on a relationship between CO2 human subject sampling. Journal of Economic Entomology 76, 327–329. Mount, G.A. and Whitney, R.W. (1984) Ultralow volume mists of chlorpyrifos from a tractor-mounted blower for area control of the lone star tick (Acari: Ixodidae). Journal of Economic Entomology 77, 1219–1223. Mount, G.A., Haile, D.G. and Daniels, E. (1997a) Simulation of blacklegged tick (Acari: Ixodidae) population dynamics and transmission of Borrelia burgdorferi. Journal of Medical Entomology 34, 461–484. Mount, G.A., Haile, D.G. and Daniels, E. (1997b) Simulation of management strategies for the blacklegged tick (Acari: Ixodidae) and the Lyme disease spirochete, Borrelia burgdorferi. Journal of Medical Entomology 34, 672–683. Mwangi, E.N., Dipeolu, O.O., Newson, R.M., Kaaya, G.P. and Hassan, S.M. (1991) Predators, parasitoids and pathogens of ticks: a review. Biocontrol Science and Technology 1, 147–156. Mwangi, E.N., Hassan, S.M., Kaaya, G.P. and Essuman, S. (1997) The impact of Ixodiphagus hookeri, a tick parasitoid, on Amblyomma variegatum (Acari: Ixodidae)

in a field trial in Kenya. Experimental and Applied Acarology 21, 117–126. Needham, G.R. and Teel, P.D. (1986) Water balance by ticks between bloodmeals. In: Sauer, J.R. and Hair, J.A. (eds) Morphology, Physiology, and Behavioral Ecology of Ticks. Ellis Horwoord, Chichester, pp. 100–164. Nicholson, M.C. and Mather, T.N. (1996) Methods for evaluating Lyme disease risks using geographic information systems and geospatial analysis. Journal of Medical Entomology 33, 711–720. Norval, R.A.I., Andrew, J.H.R. and Yunker, C.E. (1989) Pheromone-mediation of host selection in bont ticks (Amblyomma hebraeum). Science 243, 364–365. Norval, R.A.I., Sonsenshine, D.E., Allan, S.A. and Burridge, M.J. (1996) Efficacy of pheromone–acaricide-impregnated tail-tag decoys for controlling the bont tick, Amblyomma hebraeum (Acari: Ixodidae), on cattle in Zimbabwe. Experimental and Applied Acarology 20, 31–46. Ohnishi, J., Piesman, J. and de Silva, A.M. (2001) Antigenic and genetic heterogeneity of Borrelia burgdorferi populations transmitted by ticks. Proceedings of the National Academy of Sciences, USA 98, 670–675. Olsen, B., Jaenson, T.G., Noppa, L., Bunikis, J. and Bergstrom, S. (1993) A Lyme borreliosis cycle in seabirds and Ixodes uriae ticks. Nature 362, 340–342. Patrican, L.A. and Allan, S.A. (1995a) Application of desiccant and insecticidal soap treatments to control Ixodes scapularis (Acari: Ixodidae) nymphs and adults in a hyperendemic woodland site. Journal of Medical Entomology 32, 859–863. Patrican, L.A. and Allan, S.A. (1995b) Laboratory evaluation of desiccants and insecticidal soap applied to various substrates to control the deer tick Ixodes scapularis. Medical and Veterinary Entomology 9, 293–299.

Environmental Management for LB Control

Phillips, C.B., Liang, M.H., Sangha, O., Wright, E.A., Fossel, A.H., Lew, R.A., Fossel, K.K. and Shadick, N.A. (2001) Lyme disease and preventive behaviors in residents of Nantucket Island, Massachusetts. American Journal of Preventive Medicine 20, 219–224. Piesman, J. (1995) Dispersal of the Lyme disease spirochete Borrelia burgdorferi to salivary glands of feeding nymphal Ixodes scapularis (Acari: Ixodidae). Journal of Medical Entomology 32, 519–521. Pound, J.M., Miller, J.A. and LeMeilleur, C.A. (1994) Device and method for use as an aid in control of ticks and other ectoparasites on wildlife. US Patent #5,367,983. Pound, J.M., Miller, J.A., George, J.E., Oehler, D.D. and Harmel, D.E. (1996) Systemic treatment of white-tailed deer with ivermectin-medicated bait to control free-living populations of lone star ticks (Acari: Ixodidae). Journal of Medical Entomology 33, 385–394. Pound, J.M., Miller, J.A. and George, J.E. (2000) Efficacy of amitraz applied to white-tailed deer by the ‘4-poster’ topical treatment device in controlling freeliving lone star ticks (Acari: Ixodidae). Journal of Medical Entomology 37, 878–884. Rand, P.W., Lacombe, E.H., Holman, M.S., Lubelczyk, C. and Smith, R.P. Jr. (2000) Attempt to control ticks (Acari: Ixodidae) on deer on an isolated island using ivermectin-treated corn. Journal of Medical Entomology 37, 126–133. Randolph, S.E. (2000) Ticks and tick-borne disease systems in space and from space. Advances in Parasitology 47, 217–243. Randolph, S.E. and Rogers, D.J. (2000) Fragile transmission cycles of tick-borne encephalitis virus may be disrupted by predicted climate change. Proceedings of the Royal Society of London B 267, 1741–1744. Randolph, S.E. and Steele, G.M. (1985) An experimental evaluation of conventional control measures against the sheep

331

tick, Ixodes ricinus (L.) (Acari: Ixodidae). II. The dynamics of the tick–host interaction. Bulletin of Entomological Research 75, 501–518. Randolph, S.E., Green, R.M., Peacey, M.F. and Rogers, D.J. (2000) Seasonal synchrony: the key to tick-borne encephalitis foci identified by satellite data. Parasitology 121, 15–23. Ribeiro, J.M.C., Mather, T.N., Piesman, J. and Spielman, A. (1987) Dissemination and salivary delivery of Lyme disease spirochetes in vector ticks (Acari: Ixodidae). Journal of Medical Entomology 24, 201–205. Rogers, A.J. (1955) The abundance of Ixodes scapularis Say as affected by burning. The Florida Entomologist 38, 17–20. Samish, M. and Alekseev, E. (2001) Arthropods as predators of ticks (Ixodidae). Journal of Medical Entomology 38, 1–11. Samish, M. and Glazer, I. (1991) Killing ticks with parasitic nematodes of insects. Journal of Invertebrate Pathology 58, 281–282. Samish, M. and Rehacek, J. (1999) Pathogens and predators of ticks and their potential in biological control. Annual Review of Entomology 44, 159–182. Samish, M., Alekseev, E. and Glazer, I. (1999) Efficacy of entomopathogenic nematode strains against engorged Boophilus annulatus females (Acari: Ixodidae) under simulated field conditions. Journal of Medical Entomology 36, 727–732. Samish, M., Alekseev, E. and Glazer, I. (2000) Mortality rate of adult ticks due to infection by entomopathogenic nematodes. Journal of Parasitology 86, 679–684. Samsinakova, A., Kalalova, S., Daniel, M., Dusbabek, F., Honzakova, E. and Cerny, V. (1974) Entomogenous fungi associated with the tick Ixodes ricinus (L.). Folia Parasitologia (Praha) 21, 39–48. Schmidtmann, E.T. (1994) Ecologically based strategies for controlling ticks. In: Sonenshine, D.E. and Mather, T.N. (eds) Ecological Dynamics of Tick-borne

332

K.C. Stafford and U. Kitron

Zoonoses. Oxford University Press, New York, pp. 240–280. Schulze, T.L. and Harrison, R.L. (1995) Potential influence of leaf litter depth on effectiveness of granular carbaryl against subadult Ixodes scapularis (Acari: Ixodidae). Journal of Medical Entomology 32, 205–208. Schulze, T.L., McDevitt, W.M., Parkin, W.E. and Shisler, J.K. (1987) Effectiveness of two insecticides in controlling Ixodes dammini (Acari: Ixodidae) following an outbreak of Lyme disease in New Jersey. Journal of Medical Entomology 24, 420–424. Schulze, T.L., Taylor, G.C., Jordan, R.A., Bosler, E.M. and Shisler, J.K. (1991) Effectiveness of selected granular acaricide formulations in suppressing populations of Ixodes dammini (Acari: Ixodidae): short-term control of nymphs and larvae. Journal of Medical Entomology 28, 624–629. Schulze, T.L., Jordan, R.A., Vasvary, L.M., Chomsky, M.S., Shaw, D.C., Meddis, M.A., Taylor, R.C. and Piesman, J. (1994) Suppression of Ixodes scapularis (Acari: Ixodidae) nymphs in a large residential community. Journal of Medical Entomology 31, 206–211. Schulze, T.L., Jordan, R.A. and Hung, R.W. (1995) Suppression of subadult Ixodes scapularis (Acari: Ixodidae) following removal of leaf litter. Journal of Medical Entomology 32, 730–733. Schulze, T.L., Jordan, R. and Hung, R. (2000) Effects of granular carbaryl application on sympatric populations of Ixodes scapularis and Amblyomma americanum (Acari: Ixodidae) nymphs. Journal of Medical Entomology 37, 121–125. Schulze, T.L., Jordan, R.A., Hung, R.W., Taylor, R.C., Markowski, D. and Chomsky, M.S. (2001) Efficacy of granular deltamethrin against Ixodes scapularis and Amblyomma americanum (Acari: Ixodidae) nymphs. Journal of Medical Entomology 38, 344–346. Shadick, N.A., Daltroy, L.H., Phillips,

C.B., Liang, U.S. and Liang, M.H. (1997) Determinants of tick-avoidance behaviors in an endemic area for Lyme disease. American Journal of Preventive Medicine 13, 265–270. Shapiro, E.D., Gerber, M.A., Holabird, N.B., Berg, A.T., Feder, H.M., Bell, G.L., Rys, P.N. and Persing, D.H. (1992) A controlled trial of antimicrobial prophylaxis for Lyme disease after deer-tick bites. New England Journal of Medicine 327, 1769–1773. Sigal, L.H., Zahradnik, J.M., Lavin, P., Patella, S.J., Bryant, G., Haselby, R., Hilton, E., Kunkel, M., Adler-Klein, D., Doherty, T., Evans, J. and Malawista, S.E. (1998) A vaccine consisting of recombinant Borrelia burgdorferi outer-surface protein A to prevent Lyme disease. New England Journal of Medicine 339, 216–222. Smith, C.N. and Cole, M.M. (1943) Studies of parasites of the American dog tick. Journal of Economic Entomology 36, 569–572. Solberg, V.B., Neidhardt, K., Sardelis, M.R., Hoffman, F.J., Stevenson, R., Boobar, L.R. and Harlan, H.J. (1992) Field evaluation of two formulations of cyfluthrin for control of Ixodes dammini and Amblyomma americanum (Acari: Ixodidae). Journal of Medical Entomology 29, 634–638. Sonenshine, D.E. (1991) Tick pheromones. In: Biology of Ticks, Vol. 1. Oxford University Press, New York, pp. 331–369. Sonenshine, D.E. (1994) Biology of Ticks, Vol. 2. Oxford University Press, New York, 465 pp. Sonenshine, D.E. and Haines, G. (1985) A convenient method for controlling populations of the American dog tick, Dermacentor variabilis (Acari:Ixodidae) in the natural environment. Journal of Medical Entomology 22, 577–583. Sonenshine, D.E., Hamilton, J.G.C., Philips, J.S., Ellis, K.P. and Norval, R.A.I. (1992) Innovative techniques for control of tick disease vectors using pheromones. In:

Environmental Management for LB Control

Munderloh, U.G. and Kurtti, T.J. (eds) First International Conference on Tick-borne Pathogens at the Host–Vector Interface: an Agenda for Research. University of Minnesota, St Paul, pp. 308–313. Sonenshine, D.E., Allan, S.A., Norval, R.A. and Burridge, M.J. (1996) A selfmedicating applicator for control of ticks on deer. Medical and Veterinary Entomology 10, 149–154. Sood, S.K., Salzman, M.B., Johnson, B.J., Happ, C.M., Feig, K., Carmondy, L., Rubin, L.G., Hilton, E. and Piesman, J. (1997) Duration of tick attachment as a predictor of the risk of Lyme disease in an area in which Lyme disease is endemic. Journal of Infectious Diseases 175, 996–999. Stafford, K.C. III (1991) Effectiveness of carbaryl applications for the control of Ixodes dammini (Acari: Ixodidae) nymphs in an endemic residential area. Journal of Medical Entomology 28, 32–36. Stafford, K.C. III (1992a) Oviposition and larval dispersal of Ixodes dammini (Acari: Ixodidae). Journal of Medical Entomology 29, 129–132. Stafford, K.C. III (1992b) Third-year evaluation of host-targeted permethrin for the control of Ixodes dammini (Acari: Ixodidae) in southeastern Connecticut. Journal of Medical Entomology 29, 717–720. Stafford, K.C. III (1993) Reduced abundance of Ixodes scapularis (Acari: Ixodidae) with exclusion of deer by electric fencing. Journal of Medical Entomology 30, 986–996. Stafford, K.C. III (1994) Survival of immature Ixodes scapularis (Acari: Ixodidae) at different relative humidities. Journal of Medical Entomology 31, 310–314. Stafford, K.C. III (1997) Pesticide use by licensed applicators for the control of Ixodes scapularis (Acari: Ixodidae) in Connecticut. Journal of Medical Entomology 34, 552–558. Stafford, K.C. III (1998) Impact of controlled burns on the abundance of Ixodes

333

scapularis (Acari: Ixodidae). Journal of Medical Entomology 35, 510–513. Stafford, K.C. III and Magnarelli, L.A. (1993) Spatial and temporal patterns of Ixodes scapularis (Acari: Ixodidae) in south central Connecticut. Journal of Medical Entomology 30, 762–771. Stafford, K.C. III, DeNicola, A.J. and Magnarelli, L.A. (1996) Presence of Ixodiphagus hookeri (Hymenoptera: Encyrtidae) in two Connecticut populations of Ixodes scapularis (Acari: Ixodidae). Journal of Medical Entomology 33, 183–188. Stafford, K.C. III, Cartter, M.L., Magnarelli, L.A., Ertel, S. and Mshar, P.A. (1998) Temporal correlations between tick abundance and prevalence of ticks infested with Borrelia burgdorferi and increasing incidence of Lyme disease. Journal of Clinical Microbiology 36, 1240–1244. Steere, A.C., Sikand, V.K., Meurice, F., Patenti, D.L., Fikrig, E., Schoen, R.T., Nowakowski, J., Schmid, C.H., Laukamp, S., Buscarino, C., Krause, D.S. and Group, L.D.V.S. (1998) Vaccination against Lyme disease with recombinant Borrelia burgdorferi outersurface lipoprotein with adjuvant. New England Journal of Medicine 339, 209–215. Taso, J., Barbour, A.G., Luke, C.J., Fikrig, E. and Fish, D. (2001) OspA immunization decreases transmission of Borrelia burgdorferi spirochetes from infected Peromyscus leucopus mice to larval Ixodes scapularis ticks. Vector Borne and Zoonotic Diseases 1, 65–74. Tatchell, R.J. (1992) Ecology in relation to integrated tick management. Insect Science and its Application 13, 551–561. Telford, S.T. III (1993) Comments in the forum: management of Lyme disease. In: Ginsberg, H.S. (ed.) Ecology and Environmental Management of Lyme Disease. Rutgers University Press, New Brunswick, New Jersey, pp. 164–167. Uspensky, I. (1996) Tick-borne encephalitis prevention through vector control in Russia: an historical review. Review of

334

K.C. Stafford and U. Kitron

Medical and Veterinary Entomology 84, 679–689. Uspensky, I. (1999) Ticks as the main target of human tick-borne disease control: Russian practical experience and its lessons. Journal of Vector Ecology 24, 40–53. des Vignes, F., Piesman, J., Heffernan, R., Schulze, T.L., Stafford, K.C. III and Fish, D. (2001) Effect of tick removal on transmission of Borrelia burgdorferi and Ehrlichia phagocytophila by Ixodes scapularis nymphs. Journal of Infectious Disease 183, 773–778. Warshafsky, S., Nowakowski, J., Nadelman, R.B., Kamer, R.S., Peterson, S.J. and Wormser, G.P. (1996) Efficacy of antibiotic prophylaxis for prevention of Lyme disease: a meta-analysis. Journal of General Internal Medicine 11, 329–333. Weiser, J. and Rehacek, J. (1975) Nosema slovaca sp. n.: a second microsporidian of the tick Ixodes ricinus. Journal of Invertebrate Pathology 26, 411. Wikel, S.K., Bergman, D.K. and Ramachandra, R.N. (1996) Immunological-based control of bloodfeeding arthropods. In: Wikel, S.K. (ed.) The Immunology of Host–Ectoparasitic Arthropod Relationships. CAB International, Wallingford, UK, pp. 290–315. Wilkinson, P.R. (1957) The spelling of pasture in cattle tick control. Australian Journal of Agricultural Research 8, 414–423. Willadsen, P. and Jongejan, F. (1999) Immunology of the tick–host interaction and the control of ticks and tick-borne diseases. Parasitology Today 15, 258–262. Willadsen, P., Bird, P., Cobon, G.S. and Hungerford, J. (1995) Commercialisation of a recombinant vaccine against Boophilus microplus. Parasitology 110, S43–S50. Wilson, M.L. (1986) Reduced abundance of adult Ixodes dammini (Acari: Ixodidae) following destruction of vegetation. Journal of Economic Entomology 79, 693–696. Wilson, M.L. and Deblinger, R.D. (1993) Vector management to reduce the risk

of Lyme disease. In: Ginsberg, H.S. (ed.) Ecology and Environmental Management of Lyme Disease. Rutgers University Press, New Brunswick, New Jersey, pp. 126–156. Wilson, M.L., Telford, S.R. III, Piesman, J. and Spielman, A. (1988a) Reduced abundance of immature Ixodes dammini (Acari: Ixodidae) following elimination of deer. Journal of Medical Entomology 25, 224–228. Wilson, M.L., Litwin, T.S. and Gavin, T.A. (1988b) Microgeographic distribution of deer and Ixodes dammini: options for reducing the risk of Lyme disease. Annals of the New York Academy of Sciences 539, 437–440. Wormser, G.P., Nadelman, R.B., Dattwyler, R.J., Dennis, D.T., Shapiro, E.D., Steere, A.C., Rush, T.J., Rahn, D.W., Coyle, P.K., Persing, D.A., Fish, D. and Luft, B.J. (2000) Practice guidelines for the treatment of Lyme disease. Clinical Infectious Disease 31, S1–S14. Yoder, J.A. and Spielman, A. (1992) Differential capacity of larval deer ticks (Ixodes dammini) to imbibe water from subsaturated air. Journal of Insect Physiology 38, 863–869. Zhioua, E., Lebrun, R.A., Ginsberg, H.S. and Aeschlimann, A. (1995) Pathogenicity of Steinernema carpocapsae and S. glaseri (Nematoda: Steinernematidae) to Ixodes scapularis (Acari: Ixodidae). Journal of Medical Entomology 32, 900–905. Zhioua, E., Browning, M., Johnson, P.W., Ginsberg, H.S. and LeBrun, R.A. (1997) Pathogenicity of the entomopathogenic fungus Metarhizum anisopliae (Deuteromycetes) to Ixodes scapularis (Acari: Ixodidae). Journal of Parasitology 83, 815–818. Zhioua, E., Ginsberg, H.S., Humber, R.A. and Lebrun, R.A. (1999) Preliminary survey for entomopathogenic fungi associated with Ixodes scapularis (Acari: Ixodidae) in southern New York and New England, USA. Journal of Medical Entomology 36, 635–637.

Index

Page numbers in bold refer to illustrations and tables abundance, definition 36 Acari 156 acaricides 271, 302, 309, 313–317, 323 acrodermatitis chronica atrophicans (ACA) associated with Borrelia afzelii 264 characteristic sign of LB 16 description 1, 9–10 first description 1 inoculation experiment 2 Japan 211 treatment 19 see also Erythema (chronicum) migrans; manifestations; skin infection adhesin 59, 293–294 Advisory Committee on Immunization Practices (ACIP) 291 African swine fever 71 animal reservoirs 12–16 animal disease models 127–129 antibiotics 2, 12, 18–20

antibodies against Borrelia burgdorferi sensu lato 12–13 determination 16 erythema (chronicum) migrans 201 IgG 16, 255 IgM 16, 255 against OspA 57 against OspC 293 persistence 19 antigenic variation 59–60,122 argasid tick-borne borrelioses (ATBB) 176 argasid ticks 238–239 arthritis Connecticut 3, 23, 268 Europe 11 induction study 210 outbreak seasonality 251 treatment 19 treatment-resistant 56, 284 USA 11 see also joints; manifestations 335

336

Index

arthropod carrier species 153, 156–157 Asia 4, 72, 258 East Asia 211–215 enzootic cycles 105 avermectin 316 avian see bird

babesiosis 91, 92, 223, 225 bacterial adhesion, host tissues 293–294 bacterial genomics 71–72 bacteriolysis 135 bacteriophages 65–66 Bannwarth syndrome 10–11, 16 barrier hosts, definition 34–35 bears 231 behaviour feeding, Ixodes spp. 118–119, 301, 307 human 252, 304, 306 personal protection 304, 306 questing 101–103, 104, 154, 203, 254 see also human, activities birds –Ixodes pacificus relationship 235 competent reservoirs 253 complement 135, 235 hosts, Ixodes scapularis 234–235 marine 311 migrating and Borrelia garinii 161–162 migratory 70, 124, 132, 139, 209–210 passive dispersal, immature Ixodes scapularis 311 reservoir hosts 70, 131, 209, 210 seabirds 131, 162 species 158–159, 162, 183, 184–185, 208–210 see also host blebs 53–54 blood-meal identification 41–42 Boophilus spp. 310

Borrelia afzelii association with acrodermatitis chronica atrophicans 10, 264 association with mice 161 association with small rodents 163, 164 association with voles 161 distribution 72, 178, 181 host specificity 161–162 reservoir hosts 70 resistant strains 135 skin infection 126 vaccine, effectiveness 290–292 Borrelia bissettii 238 Borrelia burgdorferi sensu lato adaptive radiation 139 bacteriolysis 135 biology 47–73 cardiomyopathy 11 distribution, geographical 182, 211–215 diversity 130 DNA molecular typing 70–71 ecology 29–42, 223–242 evolution 67–69, 71–72 genome 63, 65, 98 identification 4, 187, 206–207 membrane composition 52–53 phylogenetic tree 213 prevalence in Russia 179–180 reservoir hosts 70, 163–164, 208–210 species complex 67 transformation and plating methodology 68 Borrelia burgdorferi sensu stricto distribution, geographical 178 frequency 154 heterogeneity 72 immunity against 283 nomenclature modification 224, 225 strain B31 63, 65, 98 transmission risk 240 vaccines 20

Index

Borrelia garinii association with birds 161–164 distribution, geographical 177, 181 flagellin gene (flaB) 70 frequency 154 heterogeneity 150, 204–205 host specificity 161–162 isolates from Ixodes persulcatus, Japan 205 in nervous tissue 126 migration 132–133 reservoir hosts 16, 70–71 vaccine, effectiveness 290–292 variants 177 Borrelia japonica 208, 210, 214 Borrelia lonestari 238–239 Borrelia lusitaniae 155, 163, 178 Borrelia sinica 214 Borrelia valaisiana association with birds 162, 163, 164 distribution, geographical 154–155, 178 reservoir hosts, 16,70–71

C3 120, 129 carditis 7, 8, 19 carrier hosts, definition 33–34 carrier tick, definition 32 carriers 38–40, 150–155 see also hosts; mammals; vectors case definition of Lyme borreliosis for US national surveillance 260 case reporting 259–267 cats 12, 13, 268, 315–316 cattle 12, 14–15 cattle tick (Boophilus spp.) 310 causative agents 3–4, 29, 149 see also Borrelia ceftriaxone 19, 270 Centers for Disease Control and Prevention (CDC) 290

337

chemoprophylaxis 19–20 chemotaxis 49–52, 126 chickens 15–16 children 3, 265, 282, 287–288 chromosomes 64, 65 chronic neuroborreliosis 11 chronic sclerodermic skin manifestations 10 ciprofloxacin 66 circular plasmids 65 classification of Borrelia burgdorferi 67–69, 201 clinical features 3, 5–10 see also Acrodermatitis chronica atrophicans; erythema (chronicum) migrans; manifestations clinical trials 284–288 co-evolution 72 co-feeding 34, 103, 160–161, 236–237 co-infection 39 coefficient of transmission 160 Colorado tick fever 223 complement 120, 133–134, 135, 235–237 complement regulatory acquiring surface proteins (CRASPs) 120, 133–134, 137 computer simulation model, LYMESIM 312–313, 322–323 conferences, international 4–5, 6 control 295–296, 301–324 see also prevention; vaccination cost-effectiveness analyses, per case averted 269–270 cyst formation 53–54

dark-field microscopy (DFM) 39 DDT 302 dead-end hosts 12, 30, 158 deer acaricide treatment 316 black-tailed deer 232

338

Index

deer continued borreliacidal serum 235 co-feeding transmission 161 complement 135 density 231–232, 312 exclusion 311–313 four-poster passive topical treatment 316–317 hosts for Ixodes scapularis 231–233 maintenance hosts 34, 257–258 population reduction 311–313 roe deer–Ixodes ricinus system 317 repellents sika deer 202 susceptibility to infection 15 white-tailed deer 311–313 zooprophylactic role 131 density, definition 36 developmental cycle, Ixodes spp. 100 diagnosis 16–18 disease classification 255–256 DNA detection by PCR 17 DNA molecular typing, Borrelia burgdorferi sensu lato 70–71 DNA–DNA reassociation analysis 69 dogs 12–13, 268, 282, 283, 315–316 domestic animals 12–16, 127, 184, 231, 268 see also hosts; mammals dormice 158 doxycycline 19, 20, 269–270, 271 ecology Borrelia burgdorferi sensu lato 201–242 research 29–42 risk factors 192–194 terminology 31–37 tick 225–228 ehrlichiosis, granulocytic 91, 92, 225 emergence of LB, USA 258–259 endemic regions 257–258, 259 environmental management 301–324

see also climatic conditions; habitat enzootic transmission cycles 104–105, 208–210 enzyme-linked immunosorbent assay (ELISA) 16, 261, 266, 270, 286–288, 289 epidemiology 70–71, 204, 251–271, 282–283 erythema (chronicum) migrans case definition 266 cases 251 description 7–10 ecological factors 265 Japan 211 penicillin treatment 3 rash 282 relationship with tick bite and nervous disorders 2 seasonal occurrence 267 skin lesions 16, 17 symptoms 211 treatment 19 unusual case, Wisconsin 223 see also acrodermatitis chronica atrophicans; manifestations; skin infection Eurasia 5, 8, 178 Europe arthritis 11 Borrelia afzelii, geographical distribution 72, 154 Borrelia burgdorferi s.s., geographical distribution 154, 178 Borrelia garinii, geographical distribution, 154, 177, 181 Borrelia lusitaniae, geographical distribution 155 Borrelia valaisiana, geographical distribution 154–155 clinical features 5, 8 descriptive population data 264–267 ecological correlates 258

Index

ecological research 30 enzootic maintenance cycle 104–105 seasonality of LB cases 103 serosurveys 13 therapy 19 European Union Concerted Action on Lyme borreliosis 5, 264–265 experimental reservoir hosts, definition 35 experimental vector (tick), definition 35 extracellular matrix (ECM) 58 extracellular vesicles (blebs) 53–54 feeding behaviour, Ixodes spp. 118–119, 301, 307 fibronectin-binding protein BBK32 59, 294 fipronil 315–316 flagella 49–52 focal patterns 257–258 Food and Drug Administration (FDA) 290, 291 foxes 158 fungi, entomopathogenic 319 Garin–Bujadoux–Bannwarth syndrome 10–11, 16 genetic recombination 122 genetic tools 66–67 genome and genome organization 62–67, 71–72 genospecies differences 129 differentiated with molecular methods 30 distribution, geographic 48 distribution in small mammals 185 diversity 237–238 incidence 188 new 238, 208, 214 genotypic methods, molecular characterization 69–70

339

geographic distribution see distribution goats, antibodies 15 habitat, tick 189, 228–231, 257–258, 302–304 see also environmental management hares 158, 183–184 hedgehogs 152, 158 hedgehog tick (Ixodes hexagonus) 150, 152, 158 heterogeneity 72, 150, 204–205 high-risk areas 257–258 horizontal transmission 209, 237 horses 12, 13–14 host abundance 324 barrier host, definition 34–35 bridging host, definition 35 complement 134, 235–237 dead-end 12, 30, 158 detection by ticks 99 diversity 130–131 immune status 159–160 infection detection 40 infection persistence 121–126 dissemination 119–121 immune status 159–160 infectivity, definition 36 maintenance 34, 104–105, 150–151, 257–258 maintenance hosts, definition 33 management 310 non-reservoir 160–161 of Ixodes scapularis 231–235 parasitization 258 preference, tick 226 rodent 98 species 131–132, 155–164 specificity, Borrelia spp. 131–138, 161 spirochaetes preferences 184–185 to tick transmission 126–127 tick–vertebrate cycles 104–105

340

Index

host continued tissue bacterial adhesion 293–294 tissue colonization 58–59 see also carriers; mammals; reservoir hosts; vector human clinical trials 284–288 disease agents 92 infection dynamics 255–256 infection risk 192–194 Ixodes spp. attacks 184 risk of disease 239–241 risk-associated activities and behaviours 252, 258, 266, 268, 283 tick-bite cases 201 immunoblot tests 16, 265, 270 immunofluorescence 16, 39, 50 immunoprophylaxis 20 incubation period, LB 255 infection co-infection 39 detection, vertebrate host 40 determinants 267–269 dynamics, human 255–256 early disseminated 255 early localized 255 intensity, definition 36 late persistent infection 255 mixed 182–183 prevalence, tick 38, 39 prevalence, definition 36 reservoir hosts 253–254 risk, human 192–194 skin infection, Borrelia afzelii 126 transmission to humans 254–256 Insecta 97, 156, 157 insecticides 315–316 insectivores 152, 157, 158, 183 integrated pest management (IPM) 321–323 intensity of infestation, definition 36 intervention strategies 323

intervention model 271 isolates, collection 181 ivermectin 316 Ixodes dammini 3–4, 226–227 see also Ixodes scapularis Ixodes hexagonus 150, 152, 155, 158 Ixodes jellisoni 238 Ixodes nipponensis 213 Ixodes ovatus 201–203, 204, 205, 208, 210 Ixodes pacificus abundance 103 –bird relationship 235 association with Borrelia bissettii 238 association with Columbian blacktailed deer 232 hosts 233–234 infesting lizards 103 life stages 101 principal vector 230–231 seasonal distribution of LB cases 104 Ixodes pavlovskyi 178 Ixodes persulcatus adult density 192, 194 distribution, geographical 105, 151 enzootic transmission 210 isolates from 205 landscape preference 190, 191 life cycle 203 maintenance hosts 258 preferred hosts 209 prevalence of borreliae 179–180, 204 questing activity 101–102, 104, 203 seasonal fluctuation of tick bite 202–203 transmission rates 306 vertebrate host 183–184 Ixodes ricinus abundance 101, 103 co-feeding 236–237 complement activity 135

Index

distribution, geographical 105, 151, 256–257 habitat 189 host range 310 infesting rodent 103 life cycle 152, 266–267 maintenance hosts 258 questing activity 154 rodent resistance to 159 transmission rates 306 vector in European Russia 176 vertebrate hosts 152, 155, 157–159, 164 Ixodes scapularis -borne illness 267–268 –Ixodes dammini crosses 226 abundance 101 abundance reduction 311–312 alternative mammalian host 233 bite 223 control 313–317 ecology 225–226 habitat 228–230 hosts 231–235, 236 life cycle 152, 266–267 nymphs 98, 101, 103, 240 passive dispersal 311 population genetics 227 reservoir hosts 235–236 Ixodes spinipalpis 233–234 Ixodes spp. control methods 304–307 developmental stages 99–102 distribution, geographical 150–151 feeding behaviour 118–119, 301, 307 infection 164–165 questing behaviour 101–103, 104, 154, 203, 254 reservoir hosts 186 seasonal activity 203 vaccines 295 vector competence 93, 103–104 Ixodes trianguliceps 178

341

Ixodes uriae Borrelia burgdorferi sensu lato circulation in Europe 152 DNA identified in 70 maintenance in secondary transmission cycles 150–151 Borrelia garinii association with seabirds 162 DNA amplified from 47 OspA sequences 123, 132–133 vertebrate hosts 155, 159 ixodid tick-borne borreliosis (ITBB) 176, 184, 185–186, 189–192, 193, 194

joints 11, 126, 285 see also arthritis; manifestations

Koch’s postulates 14 laboratory animal species 128–129 laboratory evidence 16–17 lagomorphs 158, 183–184 landscape management 308–310 landscape preference, Ixodes persulcatus 190, 191 late persistent infection 255 later stage manifestations 255 life cycle, tick 101, 138, 152, 203, 266–267 life stages, tick 103–104 linear chromosome 64 linear plasmids 65 lipopolysaccharide (LPS) 52–53 lipoprotein 52, 54–56, 59–60, 62, 121 lizards 34–35, 103, 104, 160, 234, 235, 236 Lyme, Connecticut, ix 1, 3, 223, 258–259 Lyme arthritis see arthritis LYMErix 281, 286, 291–292

342

Index

LYMESIM,

computer simulation model 312–313, 322–323 lymphocytoma 1, 2, 9, 16, 19, 264

maintenance hosts 34, 104–105, 150–151, 257–258 maintenance hosts, definition 33 mammals acaricide application 314–317 hosts 233 large 209 medium-sized 157, 184, 236 passive dispersal immature Ixodes scapularis 311 see also carriers; host; small mammal; vector; vertebrate host management strategies 305 manifestations canine LB 12 cardiac 11 cattle 14–15 characteristic signs 16 clinical 7–10, 149 Russian terminology 175 clinical definition 264–265 development 2–3 dynamics of infection 255–256 early neurological 10–11 equine LB 13–14 eye 7 joint 11, 12, 13 later stage 255 meningo-polyradiculoneuritis 1–2 musculoskeletal 7, 8 nervous system 7, 8 skin 7–8, 9–10 three stages of LB 5–11 see also acrodermatitis chronica atrophicans; arthritis; clinical features; erythema (chronicum) migrans; joints mean density, definition 36

membrane proteins 60–62 see also outer surface proteins (Osps); proteins membrane structure, borreliae 49–62 methodological pitfalls 29–42 MHC haplotypes 129 mice association with Borrelia afzelii 161 Borrelia burgdorferi host 310 C3H mouse strain 129 density 101, 312 immune status 159–160 mouse models 128–129, 312 reservoir competence 157 sera 235–237 white-footed mice 101, 104, 235–237, 312 see also rodents migration bird 70, 124, 132, 139, 161–162, 209–210 Borrelia spp. 72, 132–133 models animal disease models 127–129 intervention model 271 LYMESIM 312–313, 322–323 mouse models 128–129 transmission 134–138 morbidity 176, 194 morphology, borreliae 49–62 motility, borreliae 49–52 moult, ticks 39 mouse models 128–129 Museum of Borrelia, Laboratory of Vectors of Infection 181

national surveillance, USA 259–264 nematode parasites, ticks 319–320 neuroborreliosis 2, 10–11, 267 New York Times article 5 non-reservoir hosts 160–161 non-reservoir hosts, definition 34 non-vector tick, definition 31–32

Index

non-vector tick species 96 North America clinical features 5, 8 deer abundance 33 descriptive population data 258–264 ecological correlates 257–258 ecological research 30 ecology Borrelia burgdorferi sensu lato 223–242 maintenance hosts, deer 34 nymphal vs. female tick, importance 104 Osps see outer surface proteins (Osps) outer membrane proteins 54–55 outer membrane spanning (Oms) 61 outer membrane structure 52–53 outer surface proteins (Osps) characterized 49 during tick feeding 97–98 founding members of family 58 ligand for factor H 120 lipoproteins 55–58 major lipoproteins 55–58 OspA function 55–56 gene sequences 123, 124, 132–133, 213–214 intraspecific heterogeneity 150 recombinant 54 serotypes 204–205 typing system 4 vaccine 20 OspC 56–57, 125 OspD 57 OspE 58, 61, 294 OspEF-related proteins (ERP) 58 OspF 58, 294 selectively expressed 118–119 targeted 121 see also proteins

343

parasiticide 316 parasitization 183–184 pasture spelling 310 pathogen detection 37–42 pathogen-exposed hosts, definition 33 pathogen-exposed ticks, definition 32 pathogenesis 127–129 pathogenicity, Borrelia spp. 210–211 penicillin 2, 3, 19 permethrin 315 persistence, vertebrate host 121–126 personal protection 304, 306 phase contrast microscopy (PCM) 39 pheasants 16 phenotypic methods, molecular characterization 69–70 pheromones, tick 320–321 phylogenetic tree, Borrelia burgdorferi sensu lato 213 plasmids 62, 65 plasminogen system 120–121 polymerase chain reaction (PCR) technology Borrelia burgdorferi sensu lato detection method 38 diagnosis support 12–13, 14, 15 DNA detection 17 molecular characterization 69–70 RNA detection 39 preferred hosts, Ixodes persulcatus 209 pregnancy, risk for adverse outcomes 268–269 prevalence, Borrelia burgdorferi 179–180 prevalence of infection, definition 36 prevalence of infestation, definition 36 prevention 270–271, 290, 302, 308, 322 see also personal protection; vaccination; vaccine primary vectors 95, 99–102, 103–104 prophylaxis 19–20

344

Index

proteins complement regulatory acquiring surface proteins (CRASPs) 120, 133–134, 137 decorin-binding proteins (Dbps) 58–59, 294 exported plasmid protein 62 fibronectin-binding protein BBK32 59, 294 host-derived complement control proteins 120 integral membrane proteins 60–61 lipoprotein 52, 54–56, 59–60, 62, 121 outer membrane proteins 54–55 surface-exposed membrane proteins 61–62 variable lipoproteins (Vls) 59, 60 variable membrane proteins (Vmps) 59–60, 122, 124 see also outer surface proteins (Osps) pyrethrin, natural 314 pyrethroid cyfluthrin 314

qualitative terms 31–36 quality-adjusted life years (QALY) 269 quantitative terms 36–37 questing activity 101–104, 154, 203, 254

rabbit 158 rat 104, 157, 163, 233 recombinant vaccines 13, 57 relapsing fever 223 relative reservoir capacity, definition 37 relative vector capacity, definition 37 repellents 51, 306, 311 reported cases 251–252, 261, 262, 263

reservoir hosts capacity, definition 37 competence 131–132, 157, 159, 161, 235–237 definition 29–30, 32–33, 34 enzootic cycles 104 identification 40–42, 155 infection 253–254 natural 117, 127 relative reservoir capacity, definition 37 see also birds; deer; host; mammals; mice; rodents reservoir potential 160 reservoir potential, definition 37 reservoirs see reservoir hosts resistance of Borrelia burgdorferi to, host complement 134 resistant strains, Borrelia afzelii 135 restriction fragment length polymorphism analysis (RFLP) 204, 205 risk ecological factors 192–194 EUCALB 264–265 factors 267–269 high-risk areas 257–258, 282 human 239–241 infection 192–194 mapping 224, 302–304, 323–324 pregnancy, adverse outcomes 268–269 transmission 239–240 Rocky Mountain spotted fever 3, 223 rodent borrelial isolates 211 complement 135 dormice 158 enzootic transmission 210 hosts 98, 132, 232–234 infection reservoirs 253 infestation, Ixodes ricinus 103 principal reservoir hosts 117 rats 104, 157, 163, 233

Index

reservoir, habitat 257 reservoir hosts 70, 131, 163, 236, 237, 238 squirrels 158, 163 topical treatment 315 voles 159–160, 183, 186 see also host; mice; reservoir hosts Russian terminology, clinical manifestations 175

satellite imagery 303–304 scientific publications 4–5 sclerodermic skin manifestations 10 seasonality activity, ticks 102–103, 203 arthritis 3, 251 Borrelia burgdorferi sensu stricto 240–241 disease onset 263, 265, 266–267 distribution, human LB 104 erythema migrans 267 fluctuation tick bite cases 202 host-seeking in ticks 30 tick questing activity 30 101, 102–103, 203, 254 serological diagnostic testing cost 270 serosurveys 13 sheep 12, 15, 236–237 shrews 157, 158, 183 skin infection 7–8, 9–10, 16, 17, 126 see also acrodermatitis chronica atrophicans; erythema (chronicum) migrans; manifestations small mammals distribution of Borrelia burgdorferi sensu lato genospecies 185 genera 130–131 identification Borrelia burgdorferi sensu lato isolates 186–187 parasitization by larvae and nymphs 183

345

reservoir competent 132 reservoir hosts 157, 184–185 species 182, 184, 208–209 see also mammal SmithKline Beecham 286 specific host infectivity, definition 36 specific tick infectivity, definition 36 spirochaetes biology 47–73 host preferences 184–185 phenotypic and genotypic heterogeneity 150 primary bridging vectors 92 primary vector distribution 95 tick–host–spirochaete interactions diversity 129–130 squirrels 158, 163 see also rodent surface-exposed membrane proteins 61–62 surveillance statistics 264 symptoms see manifestations synthetic pyrethroid cyfluthrin 314 systematics, tick ecology 225–228

T-cell proliferation assay 17 taiga tick see Ixodes persulcatus terminology 29–42, 175 tests 16–17, 270 ticks Amblyomma americanum (lone star tick) 238–239, 283, 308 American dog tick (Dermacentor variabilis) 233 Argas spp. 92, 156 argasid ticks 238–239 biological control 318 black-legged tick see Ixodes scapularis Boophilus spp. 92, 310 bite cases 202–203 castor bean tick see Ixodes ricinus chemical control, 313–317

346

Index

ticks continued control methods 295–296, 301, 313–320 Dermacentor spp. 92, 176–177, 233 ecology 102–105, 225–228 European sheep tick see Ixodes ricinus habitat 189, 228–231, 257–258, 302–304 Haemaphysalis spp. 92, 156, 214 hard 29 hosts 100 hosts, definition 33 Hyalomma 92 infectivity, definition 36 landscape management 308–310 life stages 103–104 Lone star tick, see Amblyomma americanum microenvironment modification 308–310 natural enemies 318 non-vector, definition 32 non-vector species 96 off-host survival 101 Ornithodoros 92 pheromones 320–321 population reduction 310–313 population suppression 308 questing activity 101–103 Rhipicephalus 92, 156 saliva 98 seabird tick see Ixodes uriae soft 29 species 91–92, 94, 176, 182 survival, humidity 308 taiga tick see Ixodes persulcatus 177–183 vector–vertebrate cycles 104–105 western black-legged tick see Ixodes pacificus see also host; Ixodes; ixodid tickborne borreliosis biotopes (ITBB); vectors

tick-borne encephalitis (TBE) 91, 92, 175–176, 184, 192, 302 tick-borne zoonoses 32, 33 tick–borrelia relationships 97–99 transformation and plating methodology 68 transmembrane-spanning domains 61 transmission co-feeding 34, 103 coefficient of transmission 160 current concepts 118–139 cycle, Borrelia burgdorferi sensu lato model 134–138 enzootic transmission 208–210 horizontal 209, 237 host to tick 126–127 infection, transmission to humans 254–256 model 134–138 modes 254–255 rates 306 risk 239–240, 302–304 transovarial 105, 153–154, 239 transstadial 41, 96, 99, 105, 138, 155 definition 153 zoonotic cycles 117 see also vectors treatment 18–20, 66, 269–270 treatment-resistant Lyme arthritis 56, 284 Treponema spp. 12, 14, 71

ungulates 160, 258 see also deer; host United States of America arthritis 3, 11, 23, 268 Borrelia burgdorferi sensu stricto, heterogeneity 72 canine LB 12 emergence of LB 258–259 equine LB 13 national surveillance 259–264 reported cases 261, 262, 263

Index

risk map 224 seasonality of LB cases 103 USSR, ecological research 30

vaccination 281–296, 301, 306–308 see also prevention; control vaccine Borrelia burgdorferi sensu stricto 20 canine LB 13 equine LB 14 LB 255 recently developed 271 recombinant 13, 57 Vaccine Adverse Events Reporting System (VAERS) 290 variable lipoproteins (Vls) 59, 60 variable membrane proteins (Vmps) 59–60, 122, 124 vector associations 252–253 biology 99–102 bridging vectors 35, 92, 93, 95, 102, 103–104 capacity, definition 36–37 climatic conditions 102 competence 92–97, 103–104, 214 definition 31–32 distribution 95

347

experimental vector (tick), definition 35 identification 40–42 potential 93, 102–105 potential definition 37 relative vector capacity, definition 37 species relationships with Borrelia burgdorferi sensu lato 203–208 see also Ixodes scapularis; tick vector/reservoir competence, definition 35–36 vegetation management 308 vertebrate host see host voles 159–160, 161, 183, 186

wasp parasites, ticks 318–319 white-tailed deer 311–313 see also deer woodchip barrier 309 woodrat 237, 238

xenodiagnosis 40, 41–42

zoonoses 32, 33 zoonotic transmission cycles 117